math0411182/km.tex
1: \documentclass[12pt]{article}
2: \usepackage{amsmath}
3: \usepackage{amssymb}
4: \usepackage{amscd}
5: \usepackage{epsfig}
6: \usepackage{amsthm}
7: 
8: 
9: \def\mini{\scriptsize}
10: \def\t{\tilde}
11: \newcommand{\diam}{\operatorname{diam}}
12: \newcommand{\Isom}{\operatorname{Isom}}
13: \newcommand{\rank}{\operatorname{rank}}
14: \newcommand{\length}{\operatorname{\ul{length}_{\Delta}}}
15: \newcommand{\Length}{\operatorname{length_{\Delta}}}
16: 
17: 
18: 
19: 
20: 
21: 
22: \def\ora{\overrightarrow}
23: \def\half{\frac{1}{2}}
24: \def\no{\noindent}
25: \setlength{\parindent}{.25in} \setlength{\textwidth}{6in}
26: \setlength{\oddsidemargin}{.25in} \setlength{\evensidemargin}{.25in}
27: \setlength{\textheight}{9in} \setlength{\headheight}{0.2in}
28: \setlength{\topmargin}{-.25in}
29: \setlength{\parskip}{\smallskipamount}
30: 
31: 
32: 
33: 
34: \newtheorem{dfn}{Definition}[section]
35: \newtheorem{rem}[dfn]{Remark}
36: \newtheorem{thm}[dfn]{Theorem}
37: \newtheorem{defn}[dfn]{Definition}
38: \newtheorem{lem}[dfn]{Lemma}
39: \newtheorem{lemma}[dfn]{Lemma}
40: \newtheorem{sublem}[dfn]{Sublemma}
41: \newtheorem{prob}[dfn]{Problem}
42: \newtheorem{prop}[dfn]{Proposition}
43: \newtheorem{ass}[dfn]{Assumption}
44: \newtheorem{add}[dfn]{Addendum}
45: \newtheorem{claim}[dfn]{Claim}
46: \newtheorem{cor}[dfn]{Corollary}
47: \newtheorem{conj}[dfn]{Conjecture}
48: \newtheorem{conjecture}[dfn]{Conjecture}
49: \newtheorem{conc}[dfn]{Conclusion}
50: \newtheorem{cons}[dfn]{Consequence}
51: \newtheorem{condition}[dfn]{Condition}
52: \newtheorem{conv}[dfn]{Convention}
53: \newtheorem{ex}[dfn]{Example}
54: \newtheorem{example}[dfn]{Example}
55: \newtheorem{fact}[dfn]{Fact}
56: \newtheorem{question}[dfn]{Question}
57: \newtheorem{sit}[dfn]{Situation}
58: \newtheorem{spec}[dfn]{Special case}
59: \newtheorem{crit}[dfn]{Criterion}
60: \newtheorem{lis}[dfn]{List}
61: 
62: 
63: \def\proof{\par\medskip\noindent{\it Proof: }}
64: \def\sketch{\par\medskip\noindent{\it Sketch of proof: }}
65: \def\lra{\longrightarrow}
66: \def\Lra{\Longrightarrow}
67: \def\ra{\rightarrow}
68: \def\Ra{\Rightarrow}
69: \def\CR{\curvearrowright}
70: \def\acts{\CR}
71: \def\lh{\hookleftarrow}
72: \def\embed{\hookrightarrow}
73: \def\ti{\tilde}
74: 
75: 
76: \def\back{\backslash}
77: 
78: 
79: \def\P{{\mathbb P}}
80: \def\C{{\mathbb C}}
81: \def\R{{\mathbb R}}
82: \def\H{{\mathbb H}}
83: \def\Z{{\mathbb Z}}
84: \def\K{{\mathbb K}}
85: \def\M{{\mathcal M}}
86: \def\e{{\mathcal E}}
87: 
88: 
89: \def\O{{\mathcal O}}
90: 
91: 
92: 
93: 
94: 
95: 
96: \def\Q{{\mathbb Q}}
97: 
98: \def\B{{\mathbb B}}
99: \def\N{{\mathbb N}}
100: \def\l{{\mathbb l}}
101: \def\F{{\mathbb F}}
102: \def\g{{\mathfrak g}}
103: \def\k{{\mathfrak k}}
104: \def\p{{\mathfrak p}}
105: \def\a{{\mathfrak a}}
106: \def\eps{\epsilon}
107: \def\al{\alpha}
108: \def\be{\beta}
109: \def\ga{\gamma}
110: 
111: 
112: \def\Ga{\Gamma}
113: \def\de{\delta}
114: \def\De{\Delta}
115: 
116: 
117: \def\Del{\Delta}
118: \def\Si{\Sigma}
119: \def\si{\sigma}
120: \def\th{\theta}
121: \def\L{{\cal L}}
122: \def\la{\lambda}
123: \def\La{\Lambda}
124: \def\Om{\Omega}
125: \def\om{\omega}
126: 
127: 
128: \def\>{\rangle}
129: \def\<{\langle}
130: \def\del{\delta}
131: \def\D{\partial}
132: \def\3{\ss}
133: \def\8{\infty}
134: 
135: 
136: \def\ol{\overline}
137: \def\ul{\underline}
138: \def\ov{\overrightarrow}
139: \newcommand{\restr}{\mbox{\Large \(|\)\normalsize}}
140: \def\nbd{neighborhood~}
141: \def\mf{\mathfrak }
142: 
143: 
144: \overfullrule=0pt
145: 
146: 
147: 
148: 
149: \begin{document}
150: 
151: 
152: \title{A path model for geodesics in Euclidean buildings and its
153: applications to  representation theory}
154: \author{Misha Kapovich and John J. Millson}
155: \date{April 12, 2008}
156: \maketitle
157: 
158: 
159: \begin{abstract}
160: In this paper we give a combinatorial characterization of projections of geodesics in Euclidean buildings to
161: Weyl chambers. We apply these results to the representation theory of
162: complex reductive Lie groups and to spherical Hecke rings associated with split nonarchimedean reductive
163: Lie groups. Our main application is a generalization of the
164: saturation theorem of Knutson and Tao for $SL_n$ to other complex semisimple
165: Lie groups.
166: \end{abstract}
167: 
168: 
169: \section{Introduction}
170: 
171: Let $\ul{G}$  be a $\Q$-split reductive algebraic group defined over
172: $\Z$ and $\ul{G}^\vee$ be its Langlands' dual. In this paper we
173: continue our study (which we began in \cite{KLM3}) of the interaction
174: between the representation theory of the group
175: $G^\vee:=\ul{G}^\vee(\C)$ and geometry of the Bruhat--Tits building
176: associated with the nonarchimedean group
177: $G=\ul{G}(\K)$, where $\K$ is a complete field with discrete
178: valuation. We restrict ourselves to the case when $\K$ is a local field, in which case, algebraically speaking,
179: we will be studying the relation between
180: the representation ring of the group
181: $G^\vee$ and the {\em spherical Hecke algebra}
182: $\mathcal{H}_G$ associated with  $G$.
183: 
184: In his papers \cite{Littelmann1}, \cite{Littelmann2},
185: P.~Littelmann introduced a {\em path model} for the
186: representations  of complex reductive Lie groups $G^\vee$. The
187: Littelmann path model gives a method to compute the structure
188: constants of the representation ring of $G^\vee$ by counting
189: certain piecewise linear paths, called LS paths.
190: 
191: 
192: In this paper we define a class of piecewise-linear paths in $\De$, a Weyl chamber of the Weyl group $W$ of $G^\vee$.
193: These paths will be called {\em Hecke paths}  (see Definition \ref{Hecke}) because of their
194: connection with $\mathcal{H}_G$. We will prove that a path $p$ in $\De$ is a Hecke path if and only if $p$
195: is the projection into $\De$ of a geodesic segment in the
196: Euclidean (Bruhat-Tits) building $X$ associated with $G$.  Thus, unlike LS paths which had to be {\em invented}, the
197: Hecke paths appear very naturally as projections of geodesic
198: segments. Hecke paths are defined by eliminating one of the axioms for
199: LS paths, therefore {\em each LS path for $G^\vee$ is a Hecke path
200: for $G$}.
201: 
202: The converse relation is more subtle and is discussed later in the
203: introduction. To state our main results we need a definition of
204: $k_R$, the {\em saturation factor} of the root system $R$ of the
205: group $\ul{G}$. Let $\alpha_1,...,\alpha_l\in R$ be the simple
206: roots (corresponding to $\De$). Let $\theta$ be the highest root
207: and define positive integers $m_1,...,m_l$ by
208: $$
209: \theta = \sum_{i=1}^l m_i \alpha_i.$$
210: Then $k_R$ is the least common multiple of the numbers $m_i$, $i=1,...,l$.
211: We refer to section \ref{kfactors} for the computation of $k_R$.
212: 
213: \medskip
214: Below, $L$ is the character lattice of a maximal torus in $G^\vee$ (so that $\De\subset L\otimes \R$),
215: $Q(R^\vee)$ is the root lattice of $R^\vee$. Our main result is the following
216: theorem (see section \ref{saturationsection}), which, in a weaker
217: form, has been conjectured by S.~Kumar:
218: 
219: 
220: \begin{thm}
221: [Saturation theorem]\label{main}  Let $G^{\vee}$ and $L$ be as above. Suppose that
222: $\al,\be,\ga\in L$ are dominant characters such that $\al+\be+\ga\in
223: Q(R^\vee)$ and that there exists $N\in \N$ so that
224: $$
225: (V_{N\al}\otimes V_{N\be} \otimes V_{N\ga})^{G^\vee}\ne 0.
226: $$
227: Then for $k=k_R^2$ we have:
228: $$
229: (V_{k\al}\otimes V_{k\be} \otimes V_{k\ga})^{G^\vee}\ne 0.
230: $$
231: \end{thm}
232: 
233: Here and in what follows $V_\la$ is the irreducible representation
234: of $G^\vee$ associated with the dominant weight $\la$ of $G^\vee$.
235: Also it will be convenient to introduce integers $n_{\al,\be}(\ga^*)$,
236: the structure constants of the representation ring of the group $G^\vee$. (Here and in what follows, $\ga^*$
237: is the dominant weight contragredient to $\ga$.) Hence
238: $n_{\al,\be}(\ga^*)$ are defined by the equation
239: $$
240: V_{\alpha} \otimes V_{\be} = \bigoplus_{\ga} n_{\al,\be}(\ga^*) V_{\ga^*}.$$
241: Here the right-hand side is the decomposition of the tensor product
242: of the irreducible representations $V_{\alpha}$ and $V_{\be}$
243: into a direct sum of irreducible representations.
244: We can then formulate the above theorem as
245: $$n_{N\alpha, N\beta}(N \gamma^*)\ne 0 \Rightarrow
246: n_{k\alpha, k\beta}(k \gamma^*)\ne 0.
247: %\ \text{for all dominant weights} \ \alpha,\beta,\gamma.
248: $$
249: 
250: 
251: We will freely move back and forth between the symmetric formulation of
252: saturation in Theorem \ref{main} and the asymmetric formulation immediately
253: above.
254: 
255: 
256: 
257: As an immediate corollary of Theorem \ref{main} we obtain a new
258: proof of the saturation theorem of A.~Knutson and T.~Tao
259: \cite{KnutsonTao}:
260: 
261: \begin{cor}
262: Suppose that $R=A_l$, i.e. the semisimple part of $G^\vee$ is
263: locally isomorphic to $SL_{l+1}$. Suppose that $\al,\be,\ga$ are
264: dominant characters such that $\al+\be+\ga\in Q(R^\vee)$ and that
265: there exists $N\in \N$ so that
266: $$
267: (V_{N\al}\otimes V_{N\be} \otimes V_{N\ga})^{G^\vee}\ne 0.
268: $$
269: Then
270: $$
271: (V_{\al}\otimes V_{\be} \otimes V_{\ga})^{G^\vee}\ne 0.
272: $$
273: \end{cor}
274: 
275: Another proof of this theorem was given by H.\ Derksen and J.\ Weyman in \cite{DW}. However
276: the proofs of \cite{KnutsonTao} and \cite{DW} do not work for root systems different from $A_l$.
277: 
278: 
279: \begin{question}
280: \label{simplyconjecture}
281: Is it true that  if $G$ is a simple simply-laced group, then in Theorem \ref{main}
282: one can always take $k=1$ and in the case of non-simply laced groups the
283: smallest $k$ which suffices is $k=2$?
284: \end{question}
285: 
286: 
287: The affirmative answer to this question is supported by the odd orthogonal groups, symplectic groups and $G_2$,
288: when one gets the saturation constant $k=2$ rather than $2^2$ and $6^2$ \cite{KLM3, BK},
289: the group $Spin(8)$, when the saturation constant equals $1$
290: \cite{KKM}, as well as by a number of computer
291: experiments with the exceptional root systems and the root systems $D_l$.
292: 
293: 
294: Using the results of \cite{KLM2, KLM3} one can reformulate  Theorem
295: \ref{main} as  follows:
296: 
297: 
298: \begin{thm}
299: \label{reform} There exists a convex homogeneous cone $D_3\subset \De^3$,
300: defined by the {\em generalized triangle inequalities},
301: which depends only on the Weyl group $W$, so that the following hold:
302: 
303: 
304: 1. If a triple $(\al,\be,\ga)\in (\De\cap L)^3$ satisfies
305: $$
306: (V_{\al}\otimes V_{\be} \otimes V_{\ga})^{G^\vee}\ne 0,
307: $$
308: then $(\al,\be,\ga)\in D_3$ and
309: $$
310: \al+\be+\ga\in Q(R^\vee).
311: $$
312: 
313: 
314: 2. ``Conversely'', if $(\al,\be,\ga)\in k^2_R \cdot L^3 \cap D_3$ and
315: $\al+\be+\ga\in k^2_R \cdot Q(R^\vee)$, then
316: $$
317: (V_{\al}\otimes V_{\be} \otimes V_{\ga})^{G^\vee}\ne 0.
318: $$
319: \end{thm}
320: 
321: 
322: 
323: We now outline the steps required to prove the Theorem \ref{main}. We will need several facts about Hecke rings
324: ${\mathcal H}$. Let $\mathcal{O}$ denote the valuation ring of $\K$, then
325: $K:=\ul{G}(\mathcal{O})$ is a maximal compact subgroup in $G$.
326: The lattice $L$ defined above is the cocharacter lattice of a maximal torus $T\subset G$.
327: The Hecke ring ${\mathcal H}$, as a $\Z$-module, is freely generated by the characteristic functions
328: $\{c_\la: \la \in L \cap \Delta\}$. The
329: multiplication on ${\mathcal H}$ is defined via the convolution
330: product $\star$. Then the  structure constants $m_{\al,\be}(\ga)$ of ${\mathcal H}$
331: are defined by
332: $$
333: c_\al \star c_\be= \sum_\ga m_{\al,\be}(\ga) c_\ga.
334: $$
335: We refer the reader to  \cite{Gross},
336: \cite{KLM3} and to section \ref{heck} of this paper for more details.
337: 
338: 
339: Let $o\in X$ be the special vertex of the building $X$ which is
340: fixed by $K$. In section \ref{distances} we define the notion
341: $d_\De(x,y)$ of the $\De$-valued distance between points $x, y\in
342: X$. Given a piecewise-geodesic path $p$ in $X$ we define its
343: $\De$-length as the sum of $\De$-distances between the consecutive
344: vertices.
345: 
346: The structure constants for ${\mathcal H}$ are related to the geometry of $X$ via the following
347: 
348: 
349: \begin{thm}
350: \cite[Theorem 8.12]{KLM3}.
351: \label{Hecke=triangles}
352: The number $m_{\al,\be}(\ga)$ is equal to the product
353: of a certain positive constant  %(depending only on $G$)
354: by the  number of geodesic triangles $T\subset X$ whose vertices are {\em special vertices}
355: of $X$ with the first vertex equal to $o$ and whose $\De$-side
356: lengths are $\al, \be, \ga^*$.
357: \end{thm}
358: 
359: Theorem \ref{main} is essentially Statement 3, which follows from
360: Statements 1 and 2, of the following
361: 
362: 
363: \begin{thm}
364: \label{3->4}
365: Set $\ell:=k_R$.
366: 
367: 1. Suppose that
368: $(\al,\be,\ga)\in D_3\cap L^3$ and $\al+\be+\ga\in Q(R^\vee)$.
369: Then the structure constants $m_{\cdot,\cdot}(\cdot)$ of the Hecke ring of the group $G$ satisfy
370: $$
371: m_{\ell\al,\ell\be}(\ell\ga)\ne 0.$$
372: 
373: 2. Suppose that $\al, \be, \ga$ are dominant coweights of $\ul{G}$, such that
374: $m_{\al,\be}(\ga)\ne 0$. Then
375: 
376: $$ n_{\ell\al,\ell\be}(\ell\ga)\ne 0.$$
377: 
378: 
379: 3. As a consequence of 1 and 2 we have: Suppose that
380: $(\al,\be,\ga)\in D_3\cap L^3$ and $\al+\be+\ga\in Q(R^\vee)$.
381: Then
382: $$
383: n_{\ell^2\al,\ell^2\be}(\ell^2\ga)\ne 0.$$
384: \end{thm}
385: 
386: \begin{rem}
387: a. Part 1 of the above theorem was proven in \cite{KLM3}. Thus the point
388: of this paper is to prove Part 2 of the above theorem.
389: 
390: b. Examples in \cite{KLM3} show that both implications
391: $$
392: (\al,\be,\ga)\in D_3\cap L^3, \al+\be+\ga\in Q(R^\vee) \Rightarrow
393: m_{\al,\be}(\ga)\ne 0
394: $$
395: and
396: $$
397: m_{\al,\be}(\ga)\ne 0 \Rightarrow n_{\al,\be}(\ga)\ne 0
398: $$
399: are {\bf false} for the groups $G_2$ and $SO(5)$. Therefore the dilation by $k_R$ in both cases is necessary at least
400: for these groups.
401: \end{rem}
402: 
403: More generally, we prove (Theorem \ref{mini3->4})
404: 
405: 
406: 
407: \begin{thm}
408: \label{minus}
409: Suppose that $\al, \be, \ga\in L$ are dominant weights so that one of them
410: is the sum of minuscule weights. Then
411: $$
412: m_{\alpha, \beta}(\gamma)\ne 0 \Rightarrow
413: n_{\alpha, \beta}(\gamma)\ne 0.
414: $$
415: \end{thm}
416: 
417: The proof of Part 2 of Theorem \ref{3->4} proceeds as follows.
418: In section \ref{characterizationsection} we prove a characterization
419: theorem for folded triangles which implies:
420: 
421: 
422: \begin{thm}
423: \label{char}
424:  There exists a geodesic triangle $T\subset X$ whose
425: vertices are {\em special vertices} of $X$ and whose $\De$-side
426: lengths are $\al, \be, \ga^*$ if and only if there exists a Hecke
427: path $p: [0,1]\to \De$ of $\De$-length $\be$ so that
428: $$
429: p(0)=\al, p(1)=\ga.
430: $$
431: \end{thm}
432: 
433: 
434: 
435: The directed segments $\pi_\alpha$, the Hecke path $p$ and the (reversed) directed
436: segment $\pi_\gamma$ fit together to form a ``broken triangle'',
437: see Figure \ref{triangle.fig}. Here and in what follows $\pi_\la$ is the geodesic path
438: parameterizing the directed segment $\ov{ox}=\la$.
439: 
440: \begin{figure}[tbh]
441: \centerline{\epsfxsize=4.5in \epsfbox{triangle.eps}}
442: \caption{\sl A broken triangle.}
443: \label{triangle.fig}
444: \end{figure}
445: 
446: 
447: Then, by combining theorems \ref{Hecke=triangles} and \ref{char}, we
448: obtain
449: 
450: 
451: \begin{thm}
452: \label{comp}
453: $m_{\al,\be}(\ga)\ne 0$ if and only if there exists a Hecke path
454: $p: [0,1]\to \De$ of $\De$-length $\be$  so that
455: $$
456: p(0)=\al, p(1)=\ga.
457: $$
458: \end{thm}
459: 
460: This statement is an analogue of Littelmann's theorem which relates structure constants $n_{\al,\be}(\ga)$
461: of the representation ring with LS paths. The problem however is that not every Hecke path is an LS
462: path (even for the group $SL(3)$).
463: 
464: In order to prove Theorem \ref{main} our ``path model'' for the nonvanishing of
465: the Hecke structure constants must be generalized to a model
466: where Hecke paths are replaced by {\em generalized Hecke
467: paths} (we also have to replace the LS-paths by {\em generalized LS paths}). More precisely,
468: begin with a geodesic triangle $[x,y,z]\subset X$ whose vertices are special vertices of $X$ and whose $\De$-side lengths
469: are $\be, \ga^*, \al$. Now replace the geodesic segment $\ol{xy}$
470: with a certain piecewise-geodesic path $\t{p}\subset X$
471: connecting $x$ and $y$ in $X$, and
472: which is contained in the 1-skeleton of a single apartment in $X$. We then show that the projection $p$ of $\t{p}$
473: to $\De$ is a {\em generalized Hecke path}. The $\De$-length of the path $p$ still equals $d_\De(x,y)=\be$.
474: 
475: The generalized Hecke paths have the advantage over Hecke
476: paths that  {\em their break points occur only at vertices of the building}.
477: By using this observation we obtain
478: 
479: \begin{thm}\label{dilate}
480: If $p$ is a generalized Hecke path with $p(0)=0$, then $k_R \cdot p$, the image
481: of $p$ under dilation by $k_R$, is a generalized LS path of $\De$-length equal to $k_R \Length(p)$.
482: \end{thm}
483: 
484: Lastly, for {\em generalized LS paths} we prove, by modifying slightly Littelmann's arguments, the following:
485: 
486: \begin{thm}
487: \label{Li}
488: $n_{\al,\be}(\ga)\ne 0$ if and only if there exists a generalized LS path $q$ with $\De$-length $\be$ so that
489: $$
490: \al+q(1)=\ga
491: $$
492: and the concatenation $\pi_\al*q$ is contained in $\De$.
493: \end{thm}
494: 
495: 
496: Theorem \ref{main} is now obtained in 2 steps, the first of which is contained in \cite{KLM3} and the second
497: is at the heart of the present paper:
498: 
499: Step 1. It was shown in \cite[Theorem 9.17]{KLM3} that
500: $$
501: n_{N\al,N\be}(N\ga)\ne 0 \Rightarrow m_{N\al,N\be}(N\ga)\ne 0.
502: $$
503: Therefore, the vector $(N\al, N\be, N\ga^*)$ belongs to the homogeneous cone $D_3$.
504: Hence, by Part 1 of Theorem \ref{3->4}
505: we conclude that $m_{\ell\al,\ell\be}(\ell\ga)\ne 0$.
506: 
507: Step 2. $m_{\ell\al,\ell\be}(\ell\ga)\ne 0$ implies existence of a geodesic triangle
508: $[z, x, y]\subset X$ with the special vertices and $\De$-side lengths $\ell\al, \ell\be, \ell\ga^*$.
509: Then by projecting the corresponding path $\t{p}$ to $\De$ and dilating it by $k_R$ we obtain
510: a path $k\al+q(t)$ in $\De$ connecting $k\al$ to $k\ga^*$, where $q(t)$ is a generalized LS path of
511: $\De$-length $k\be$. (Recall that $k=\ell^2$.) Therefore, by appealing to Theorem \ref{Li}, we see that
512: $$
513: n_{k\al,k\be}(k\ga)\ne 0
514: $$
515: which concludes the proof.
516: 
517: 
518: 
519: 
520: 
521: \medskip
522: This paper is organized as follows. Preliminary material is
523: discussed in section \ref{prelim}, where we review the concepts of
524: Coxeter complexes, buildings and piecewise-linear paths in buildings, as well as
525: the {\em generalized distances} in buildings. In section
526: \ref{sectionchains} we define the notion of {\em chains}, which is
527: essentially due to Littelmann. This concept allows one to define
528: both LS paths and Hecke paths, as well as to relate Hecke paths and
529: foldings of geodesic paths, which is discussed in the next section.
530: 
531: 
532: The main technical tool of this paper is the concept of {\em
533: folding} of geodesics in a building into an apartment, which is done
534: via retraction of the entire building to an apartment or chamber.
535: \footnote{A similar idea is used by S. Gaussent and P. Littelmann in
536: \cite{GL}, where they fold galleries in a building to galleries in
537: an apartment.} Properties of foldings are discussed in section
538: \ref{sectionfolding}. In section \ref{euclideanfolding} we prove
539: that the image of each geodesic segment in $X$ under {\em folding}
540: $f: X\to \De$ of $X$ to a Weyl chamber, is a Hecke path, Theorem \ref{chaincondition}.
541: We then prove a partial converse to this result, i.e. that each
542: Hecke path which satisfies the {\em simple chain condition} can be
543: unfolded in $X$. We also find a necessary and sufficient condition
544: for  unfolding of a path $p$ which is {\em local}, i.e. it depends
545: only on germs of the path $p$ at its break-points.
546: 
547: 
548: 
549: 
550: In section \ref{LSpolygons} we review Littelmann's
551: path model for the representation theory of complex semisimple Lie
552: groups, in particular we discuss LS paths and generalized LS paths
553: as well as raising and lowering operators.
554: 
555: 
556: We use {\em approximation of LS paths by paths satisfying the simple
557: chain condition} to unfold LS paths in $X$, Theorem \ref{LSfolded}
558: in section \ref{unfoldingLS}. Although there are Hecke paths which
559: are not LS paths, since the unfolding condition is local, by restricting
560: the root system we reduce the general unfolding problem to the case
561: of the LS paths. We thus establish that a path in $\De$ is
562: unfoldable in $X$ it and only if it is a Hecke path, this is done in
563: section \ref{characterizationsection}, Theorem
564: \ref{chain-unfolding}. The reader interested only in the proof of
565: the saturation theorem can omit this section.
566: 
567: 
568: 
569: We prove the saturation theorem in section \ref{saturationsection},
570: Corollary \ref{corQ3->Q4}. As we stated above, the idea of the proof is to replace Hecke
571: paths with piecewise-linear paths contained in the 1-skeleton of the Euclidean
572: Coxeter complex. We show that dilation by $k_R$ of such a path
573: results in a {\em generalized LS path} which in turn suffices for
574: finding nonzero invariant vectors in triple tensor products.
575: 
576: \paragraph{Hecke paths and folded galleries.} It is interesting to ask what the relationship is between our paper
577: using Hecke paths and the results of S. Gaussent and P.
578: Littelmann in \cite{GL} and others using positively folded galleries:
579: 
580: (a) {\em  Hecke paths} correspond to the {\em positively folded galleries} defined in \cite{GL}.
581: However this correspondence can be established only {\bf after} the folding--unfolding results of the present
582: paper are proven.
583: Therefore it appears that
584: one cannot prove our results characterizing Hecke paths as projections of geodesic segments
585: in building using the results of \cite{GL} and vice versa.
586: 
587: 
588: (b) In \cite{Schwer}, C. Schwer has used the gallery approach to
589: compute the Hecke structure constants $m_{\alpha,\beta}(\gamma)$.
590: In an earlier version of this paper we applied our theory to
591: compute some Hecke structure coefficients. Thus  both Hecke path
592: and gallery models can be used  to compute the structure constants
593: of the spherical Hecke algebra.
594: 
595: (c) One can prove that
596: $$n_{\al,\be}(\ga)\ne 0\Rightarrow m_{\al,\be}(\ga)\ne 0$$
597: \cite[Theorem 9.17]{KLM3} using both Hecke paths (as it is done in section \ref{characterizationsection})
598: and positively folded galleries.
599: 
600: (d) There does not  seem to be a direct way to carry over our
601: proof of the main Theorem \ref{main} or part 2 of Theorem
602: \ref{3->4} to a proof using positively folded galleries. It is
603: clear from the above, that one of the critical steps in our
604: argument is provided by Theorem \ref{dilate} stating that dilation
605: by $k_R$ converts generalized Hecke paths to generalized LS paths.
606: But there does not appear to be a way to ``stretch'' a folded
607: gallery by the factor $k_R$ --- or, more generally, to produce an
608: LS gallery from a positively folded gallery. This is an advantage
609: of the Hecke paths, which are {\em geometric} objects, over
610: positively folded galleries, which are {\em combinatorial}
611: objects. Even when $k_R=1$, there is no obvious (at least to us)
612: reason why existence of a positively folded gallery would imply
613: existence of an LS gallery. We note that if $k_R \neq 1$ then the
614: existence of a positively folded gallery does not imply the
615: existence of an LS gallery, there are counterexamples for $SO(5)$
616: and $G_2$.
617: 
618: \medskip
619: {\bf Acknowledgments.} This paper grew out of our joint work
620: \cite{KLM1, KLM2, KLM3} with Bernhard Leeb. We are grateful to him
621: for all his ideas that he contributed to those papers, especially
622: the idea of folding that came out of his contribution to
623: \cite{KLM3}. We are also grateful to Tom Haines for valuable discussions. We thank the referee for his/her remarks.
624: Part of the work on this paper was done when the first author was
625: visiting the Max Plank Institute (Bonn), during this work he was also supported by
626: the NSF grant DMS-02-03045 and DMS-04-05180. The second author was
627: supported by the NSF grant DMS-01-04006 and DMS-04-05606. Together, the authors were suported by the NSF grant
628: DMS-05-54349. The authors gratefully acknowledge support of these institutions.
629: 
630: 
631: \tableofcontents
632: 
633: \section{Definition and notation}
634: \label{prelim}
635: 
636: 
637: 
638: 
639: 
640: 
641: 
642: 
643: \subsection{Root systems and Coxeter complexes}
644: 
645: 
646: A (discrete, nonnegatively curved) {\em Coxeter complex} is a pair
647: $(A,W)$, where $A$ is either a Euclidean space or the unit sphere
648: and $W$ is a discrete reflection group acting on $A$. The {\em rank} of
649: the Coxeter complex is the dimension of $A$. The group $W$ is called {\em Weyl group} of the Coxeter complex.
650: It is called an {\em affine Weyl group} if $A$ is a Euclidean space.
651: 
652: 
653: %\begin{rem}
654: %There are also Euclidean Coxeter complexes which are nondiscrete (i.e. the
655: %group $W$ is nondiscrete, but has discrete linear part), as well as hyperbolic Coxeter complexes,
656: %where one uses reflection groups of isometries of the hyperbolic
657: %space. However we do not need them in the present paper.
658: %\end{rem}
659: 
660: An isomorphism of Coxeter complexes $(A,W), (A',W')$ is an isometry $\iota: A\to A'$ so that
661: $$
662: \iota W \iota^{-1}=W'.
663: $$
664: 
665: {\em Walls} in
666: $(A,W)$ are fixed point sets of reflections $\tau\in W$. A point
667: $x\in A$ is called {\em regular} if it does not belong to any wall
668: and {\em singular} otherwise. The closure of each connected
669: component of the set of regular points is called an {\em alcove} in
670: the Euclidean case and a {\em chamber} in the spherical case.
671: 
672: In the case when $W$ acts cocompactly on $A$, each alcove (resp. chamber) is a product (resp. join) of simplices
673: and $(A,W)$ determines structure of a {\em polysimplicial complex} on $A$. In general
674: there exists a totally-geodesic subspace
675: $A'\subset A$ which is $W$--invariant and such that $A'/W$ is compact.
676: Therefore each alcove in $A$ is a product
677: of simplices and a Euclidean subspace in $A$. Thus much of the discussion
678:  of Euclidean Coxeter complexes can be reduced to the case when $A/W$ is compact.
679: 
680: \begin{rem}
681: \label{tri}
682: A  triangulation of a fundamental alcove (chamber) in $A$ determines a
683: $W$--invariant simplicial complex on $A$. Thus
684: we can always think of $A$ as a simplicial complex.
685: \end{rem}
686: 
687: A {\em half-apartment} in $A$ is the closure of a
688: connected component of $A\setminus H$, where $H$ is a wall in $A$.
689: 
690: 
691: ``Most'' Coxeter complexes are associated with root systems as we
692: describe below. Suppose that $R$ is a root system on a vector space
693: $V$ (i.e. each element $\al\in R$ is a linear functional on $V$).
694: The {\em rank} of $R$ is the number of simple roots in $R$, i.e. is the rank of the free abelian
695: subgroup in $V^*$ generated by $R$.
696: Let $A$ denote the Euclidean affine space corresponding to $V$. This
697: data defines a finite Coxeter group $W_{sph}$, which is a reflection
698: group generated by reflections in the hyperplanes $H_{\al}=\{ x:
699: \al(x)=0\}$. {\em Weyl chambers} of $W_{sph}$ are closures of the
700: connected components of the complement to
701: $$
702: \cup_{\al\in R} H_\al.
703: $$
704: In what follows we fix a {\em positive Weyl chamber} $\De$, it
705: determines the subset of positive roots  $R^+\subset R$ and of {\em
706: simple roots} $\Phi\subset R^+$. We also have the group of coweights
707: $P(R^\vee)$ associated with $R$:
708: $$
709: \la\in P(R^\vee) \iff \forall \al\in R, \quad \al(\la)\in \Z.
710: $$
711: Let $W_{aff}$ denote the {\em affine Coxeter group} determined by
712: the above data, this group is generated by reflections in the
713: hyperplanes ({\em affine walls})
714: $$
715: H_{\al,t}=\{\al(x)=  t\}, t\in \Z.$$
716: 
717: Given a vector $\ga\in \De$, we define the {\em contragredient vector} $\ga^*$
718: as $w_0(-\ga)$, where $w_0$ is the longest element of $W$.
719: In other words, $\ga^*$ is the intersection of the $W$-orbit
720: $W\cdot (-\ga)$ with the chamber $\De$.
721: 
722: 
723: The translation subgroup in $W_{aff}$ is the {\em coroot lattice}
724: $Q(R^\vee)$,  it is generated by the coroots $\al^\vee, \al\in R$.
725: The group  $P(R^\vee)$ is the normalizer of $W_{aff}$ in the group
726: of translations $V$. Note that $V/P(R^\vee)$ is compact. The
727: dimension of this quotient is the same as the dimension of $V$
728: provided that $rank(R)$ equals the dimension of $V$.
729: 
730: 
731:  The {\em special vertices} of a Euclidean Coxeter complex are the
732: points whose stabilizer in $W_{aff}$ is isomorphic to $W_{sph}$.
733: Equivalently, they are the points in the $P(R^\vee)$-orbit of the
734: origin.
735: 
736: 
737: \begin{rem}
738: If $A/W_{aff}$ is compact, the  vertices of $(A,W_{aff})$ are the
739: vertices of the polysimplicial complex determined by $W_{aff}$.
740: \end{rem}
741: 
742: Let $A^{(0)}$ denote the {\em vertex
743: set} of $(A, W_{aff})$, which consists of points of maximal
744: intersection of walls in $A$. If $R$ spans $V^*$, the set $A^{(0)}$
745: equals the vertex set of the polysimplicial complex in $A$ defined
746: by tessellation of $A$ via alcoves of $W_{aff}$.
747: 
748: Given a Coxeter complex $(A,W)$ and a point $x\in A$ we define a new
749: Coxeter complex $(S_x,W_x)$ where $S_x$ is the unit tangent sphere
750: at $x$ and $W_x$ is the stabilizer of $x$ in $W$.
751: 
752: 
753: For a nonzero vector $\nu\in V$ we let $\bar{\nu}:= \nu/|\nu|$
754: denote the {\em normalization} of $\nu$. We define {\em rational}
755: elements of the unit sphere $S$ to be the unit vectors of the form
756: $$
757: \eta=\bar\nu, \quad \nu\in P(R^\vee).
758: $$
759: The next lemma follows immediately from compactness of
760: $V/P(R^\vee)$:
761: 
762: 
763: \begin{lem}
764: \label{density} Rational points are dense in $S$.
765: \end{lem}
766: 
767: 
768: Suppose that $(A, W)$ is a Euclidean Coxeter complex. A {\em
769: dilation} of $(A, W)$ is a dilation $h$ (i.e. a composition of
770: translation and similarity $v\mapsto \la v, \la>0$) in the affine
771: space $A$ so that
772: $$
773: h W h^{-1} \subset W.
774: $$
775: We let $Dil(A, W)$ denote the semigroup of dilations of the complex
776: $(A, W)$. We will refer to the number $\la$ as the {\em conformal
777: factor} of the dilation $h$.
778: 
779: 
780: Given a point $x\in A$ and a dilation $h\in Dil(A,W)$, we can define
781: a new spherical Coxeter complex $(S_x, W'_x)$ on the unit tangent
782: sphere $S_x$ at $x$ via pull-back
783: $$
784: W'_x:= h^*(W_{t(x)}),
785: $$
786: where $W_{h(x)}$ is the stabilizer of $h(x)$ in $W$.
787: 
788: 
789: 
790: 
791: \begin{defn}
792: \label{relation} Suppose that $W$ is a finite Coxeter group acting
793: on a vector space $V$. Define a (nontransitive) relation $\sim_W$ on
794: $V\setminus \{0\}$ by
795: $$
796: \mu \sim_W \nu \iff
797: $$
798: \centerline{$\mu, \nu$ belong to the same Weyl chamber of $W$.}
799: 
800: 
801: \noindent We will frequently omit the subscript $W$ in this
802: notation.
803: \end{defn}
804: 
805: 
806: 
807: 
808: \begin{defn}
809: \cite[page 514]{Littelmann2}
810: \label{porder} We say that
811: nonzero vectors $\nu, \mu\in V$ satisfy $\nu \rhd_W \mu$ (for short,
812: $\nu \rhd \mu$) if for each positive root $\al$,
813: $$
814: \al(\nu)< 0 \Rightarrow \al(\mu)\leq 0.
815: $$
816: \end{defn}
817: 
818: \begin{lem}
819: \label{difference}
820: Suppose that $\nu, \mu\in P(R^\vee)$, $w\in W=W_{aff}$ is such that $w(\nu)=\mu$. Then
821: $$
822: \mu-\nu\in Q(R^\vee).
823: $$
824: \end{lem}
825: \proof The mapping $w$ is a composition of reflections $\tau_i\in W$.
826: Therefore it suffices to prove the assertion in case when $w$
827: is a reflection $\tau$. This reflection is a composition of
828: a translation $t$  and a reflection $\si\in W_o$.
829: The translation $t$ belongs to the translation subgroup $Q(R^\vee)$ of $W$,
830: therefore it suffices to consider the case when $\tau=\si\in W_o$.
831: Then $\tau=\tau_\be$, where $\be$ is a root and we have
832: $$
833: \mu-\nu= -\be(\nu)\be^\vee.
834: $$
835: Since $\be(\nu)\in \Z$ and $\be^\vee\in Q(R^\vee)$, the assertion follows. \qed
836: 
837: 
838: 
839: 
840: 
841: \subsection{Paths}
842: \label{paths}
843: 
844: Suppose that $A, V, W_{aff}$, etc., are as in the previous section.
845: 
846: 
847: Let $\tilde{\mathcal P}$ denote the set of all piecewise-linear paths $p: [a,b]\to
848: V$.  We will be identifying paths that differ by
849: orientation-preserving re-parameterizations $[a,b]\to [a',b']$.
850: Accordingly, we will always (re)parameterize a piecewise-linear path with
851: constant speed. We let $p'_-(t), p'_+(t)$ denote the derivatives of the function $p$
852: from the left and from the right. The space $\t{\mathcal P}$ will be
853: given the topology of uniform convergence.
854: 
855: 
856: If $p, q: [0,1]\to A$ are piecewise-linear paths in a simplicial complex such that
857: $p(1)=q(0)$, we define their {\em composition} $r=p\cup q$ by
858: $$
859: r(t)=\left\{
860: \begin{array}{c}
861: p(t), ~~t\in [0,1],\\ q(t-1), ~~t\in [1, 2].
862:  \end{array}\right.
863: $$
864: Let ${\mathcal P}\subset \tilde{\mathcal P}$ denote the set of paths
865: $p: [0,1]\to V$ such that $p(0)=0$. Given a path $p\in {\mathcal P}$
866: we let $p^*\in {\mathcal P}$ denote the {\em reverse} path
867: $$
868: p^*(t)= p(1-t)-p(1).
869: $$
870: For a vector $\la\in V$ define a
871: geodesic path $\pi_\la\in {\mathcal P}$ by
872: $$
873: \pi_\la(t)=t\la, \quad t\in [0,1].
874: $$
875: Given two paths $p_1, p_2\in {\mathcal P}$ define their {\em
876: concatenation} $p=p_1* p_2$ by
877: $$
878: p(t)= \left\{
879: \begin{array}{c}
880: p_1(2t), \quad t\in [0, 1/2],\\ p_1(1)+ p_2(2t-1), \quad t\in [1/2,
881: 1].
882:  \end{array}\right.
883: $$
884: 
885: \medskip
886: Suppose that $p\in {\mathcal P}$ and $J=[a,b]$ is nondegenerate
887: subinterval in $I=[0,1]$. We will use the notation  $p|J\in
888: \tilde{\mathcal P}$ to denote the function-theoretic restriction of
889: $p$ to $[a,b]$. We will use the notation $p|_J$ to denote the path
890: in ${\mathcal P}$ obtained from $p|J$ by pre-composing $p|J$ with an
891: increasing linear bijection $\ell: I\to J$ and post-composing it
892: with the translation by the vector $-p(a)$.
893: 
894: \medskip
895: Fix a positive Weyl chamber $\De\subset V$; this determines the set of positive roots $R^+\subset R$,
896: the set of simple roots $\Phi\subset R^+$. We define the
897: subset ${\mathcal P}^+\subset {\mathcal P}$ consisting of the paths
898: whose image is contained in $\De$.
899: 
900: For a path $p\in {\mathcal P}$ and a positive root $\al\in R^+$ define the
901: {\em height} function
902: $$
903: h_\al(t)= \al(p(t))
904: $$
905: on $[0,1]$. Let $m_\al=m_\al(p)\in \R$ denote the minimum of
906: $h_\al$. Clearly $m _\al(p)\le 0$ for all $p\in {\mathcal P}$.
907:  We define the set of ``integral paths''
908: $$
909: {\mathcal P}_\Z:= \{ p\in {\mathcal P}: \forall \al\in \Phi,
910: m_\al(p)\in \Z\}.
911: $$
912: More restrictively, we define the set ${\mathcal P}_{\Z,loc}$ of paths $p\in {\mathcal
913: P}$  which satisfy the following {\em local integrality condition}:
914: 
915: 
916: For each simple root $\al\in \Phi$ the function $h_\al$  takes
917: integer values at the points of local minima.
918: 
919: 
920: 
921: 
922: 
923: 
924: 
925: \subsection{The saturation factors associated to a root system}
926: \label{kfactors}
927: 
928: 
929: In this section we define and compute {\em saturation factors}
930: associated with root systems.  Let $o\in A$ be a
931: special vertex, which we will identify with
932: $0\in V$.
933: 
934: 
935: \begin{dfn}
936: We define the {\em saturation factor} $k_R$ for the root system $R$
937: to be the least natural number $k$ such that $k\cdot A^{(0)}\subset
938: P(R^\vee)\cdot o$. The numbers $k_R$ for the irreducible root
939: systems are listed in the table (\ref{ta}).
940: \end{dfn}
941: 
942: 
943: Note that the condition that $k\cdot A^{(0)}\subset P(R^\vee)\cdot
944: o$ is equivalent to that each point of $k\cdot A^{(0)}$ is a special
945: vertex.
946: 
947: 
948: 
949: 
950: Below we explain how to compute the saturation factors $k_R$
951: following \cite{KLM3}. First of all, it is clear that if the root
952: system $R$ is reducible and $R_1,...,R_s$ are its irreducible
953: components, then $k_R= LCM(k_{R_1},...,k_{R_s})$, where  $LCM$
954: stands for the {\em
955:   least common multiple}. Henceforth we can assume that the system
956: $R$ is reduced, irreducible and has rank $n=\dim(V)$. Then the
957: affine Coxeter group $W_{aff}$ acts cocompactly on $A$ and its
958: fundamental domain (a {\em Weyl alcove}) is a simplex.
959: 
960: 
961: Let $\{\al_1,...,\al_{\ell}\}$ be the collection of simple roots in $R$
962: (corresponding to the positive Weyl chamber $\De$) and $\theta$ be
963: the highest root. Then
964: 
965: 
966: \begin{equation}
967: \label{highestroot} \theta= \sum_{i=1}^{\ell} m_i \al_i.
968: \end{equation}
969: 
970: 
971: We have
972: 
973: 
974: \begin{lem}
975: \cite[Section 2]{KLM3}
976: \label{kexists} $k_R=
977: LCM(m_1,...,m_n)$.
978: \end{lem}
979: 
980: 
981: Below is the list of saturation factors:
982: 
983: 
984: \begin{equation}
985: \label{ta}
986: \begin{array}{|c|c|c|}
987: \hline ~ & ~ & ~  \\ \hbox{Root system} & \theta &  k_R \\ \hline
988: A_\ell & \al_1+...+\al_\ell & 1\\ \hline B_\ell & \al_1+
989: 2\al_2+...+2\al_\ell & 2\\ \hline C_\ell & 2\al_1+
990: 2\al_2+...+2\al_{\ell-1}+\al_\ell & 2\\ \hline D_\ell & \al_1+
991: \al_2+\al_3+ 2\al_4+...+2\al_\ell & 2 \\ \hline G_2 & 3\al_1+2\al_2&
992: 6 \\ \hline F_4 & 2\al_1+ 3\al_2+4\al_3 +2\al_4 &  12\\ \hline E_6 &
993: \al_1+\al_2+ 2\al_3+2\al_4+2\al_5+3\al_6 &  6 \\ \hline E_7 & \al_1+
994: 2\al_2+2\al_3+2\al_4+3\al_5+&  12 \\ ~ & +3\al_6+ 4\al_7 & ~ \\
995: \hline E_8 & 2\al_1+2\al_2+3\al_3+3\al_4+ 4\al_5+ &  60\\ ~&
996: +4\al_6+ 5\al_7 +6\al_8 & ~  \\ \hline
997: \end{array}
998: \end{equation}
999: 
1000: 
1001: \subsection{Buildings}
1002: 
1003: 
1004: Our discussion of buildings follows \cite{KleinerLeeb}. We refer the reader
1005: to \cite{Brown}, \cite{Ronan}, \cite{Rousseau} for the more combinatorial discussion.
1006: 
1007: 
1008: Fix a spherical or Euclidean (discrete) Coxeter complex $(A,W)$, where $A$ is a
1009: Euclidean space $E$ or a unit sphere $S$ and $W=W_{aff}$ or $W=W_{sph}$
1010: is a discrete Euclidean or a spherical Coxeter group acting on $A$.
1011: 
1012: A metric space $Z$ is called {\em geodesic} if any pair of points $x, y$ in $Z$ can be connected by a
1013: geodesic segment $\ol{xy}$.
1014: 
1015: 
1016: Let $Z$ be a metric space. A {\em geometric structure} on $Z$ {\em
1017: modeled on} $(A,W)$ consists of an atlas of isometric embeddings
1018: $\varphi:A\embed Z$ satisfying the following compatibility
1019: condition: For any two charts $\varphi_1$ and $\varphi_2$, the
1020: transition map $\varphi_2^{-1}\circ\varphi_1$ is the restriction of
1021: an isometry in $W$. The charts and their images,
1022: $\varphi(A)=a\subset Z$, are called {\em apartments}. We will
1023: sometimes refer to $A$ as the {\em model apartment}. We will require
1024: that there are {\em plenty of apartments} in the sense that any two
1025: points in $Z$ lie in a common apartment. All $W$-invariant notions
1026: introduced for the Coxeter complex $(A, W)$, such as rank, walls, singular
1027: subspaces, chambers etc., carry over to geometries modeled on $(A,
1028: W)$. If $a, a'\subset X$ are alcoves (in the Euclidean case) or chambers
1029: (in the spherical case) then there exists an apartment $A'\subset X$ containing $a\cup a'$:
1030: Just take regular points $x\in a, x'\in a'$ and an apartment $A'$ passing through $x$ and $x'$.
1031: 
1032: 
1033: 
1034: A geodesic metric space $Z$ is said to be a $CAT(\kappa)$-space if
1035: {\em geodesic triangles in $Z$ are ``thinner'' than geodesic
1036: triangles in a simply-connected complete surface of the constant
1037: curvature $\kappa$}. We refer the reader to \cite{Ballmann}  for the
1038: precise definition. Suppose that $Z$ is a (non-geodesic) metric space with the discrete metric:
1039: $$
1040: d(x,y)=\pi \iff x\ne y.
1041: $$
1042: We will regard such a space as a $CAT(1)$ space as well.
1043: 
1044: \begin{dfn}
1045: A {\em spherical building} is a $CAT(1)$-space modeled on a spherical
1046: Coxeter complex.
1047: \end{dfn}
1048: 
1049: 
1050: Spherical buildings have a natural structure of polysimplicial
1051: piecewise spherical complexes. We prefer the geometric to the
1052: combinatorial view point because it appears to be more appropriate
1053: in the context of this paper.
1054: 
1055: 
1056: 
1057: 
1058: \begin{defn}
1059: A Euclidean building is a $CAT(0)$-space modeled on a (discrete)
1060: Euclidean Coxeter complex.
1061: \end{defn}
1062: 
1063: 
1064: 
1065: 
1066: A building is called {\em thick} if every wall is an intersection of
1067: apartments. A non-thick building can always be equipped with a
1068: natural structure of a thick building by reducing the Coxeter group  \cite{KleinerLeeb}.
1069: 
1070: \medskip
1071: Let $\K$ be a local field with a (discrete) valuation $\nu$ and valuation ring $\O$.
1072: Throughout this paper, we will be working with $\Q$-split reductive algebraic groups (i.e. Chevalley groups) $\ul{G}$ over $\Z$.
1073: We refer the reader to \cite{Borel, Demazure} for a detailed discussion of such groups. A reader
1074: can think of $\K=\Q_p$ (the $p$-adic numbers) and a {\em classical} Chevalley group $\ul{G}$,
1075: e.g. $GL(n)$ or $Sp(n)$, in which case  Bruhat-Tits buildings
1076: below can be described rather explicitely, see e.g. \cite{Garrett}.
1077: 
1078: Given a group $\ul{G}$ as above, we get a {\em nonarchimedean Lie
1079: group} $G=\ul{G}(\K)$, to which we can associate a Euclidean building (a
1080: Bruhat-Tits building) $X=X_G$. We refer the reader to \cite{BT},
1081: \cite{KLM3} and \cite{Rousseau} for more detailed discussion of the
1082: properties of $X$. Here we only recall that:
1083: 
1084: 
1085: 1. $X$ is thick and locally compact.
1086: 
1087: 
1088: 2. $X$ is modeled on a Euclidean Coxeter complex $(A,W_{aff})$ whose
1089: dimension equals the rank of $\ul{G}$, and the root system is
1090: isomorphic to the root system of $\ul{G}$.
1091: 
1092: 3. $X$ contains a special vertex $o$ whose stabilizer in $G$ is $\ul{G}(\O)$.
1093: 
1094: 
1095: \begin{ex}
1096: Let $X$ be a (discrete) Euclidean building, consider the  {\em
1097: spaces of directions} $\Si_x X$. We will think of this space as the
1098: space of germs of non-constant geodesic segments $\ol{xy}\subset X$.
1099: As a polysimplicial complex $\Si_xX$ is just the link of the point
1100: $x\in X$. The space of directions has the structure of a spherical
1101: building modeled on $(S,W_{sph})$, which is thick if and only if $x$
1102: is a special vertex of $X$  \cite{KleinerLeeb}. The same applies
1103: in the case when $X$ is a spherical building.
1104: \end{ex}
1105: 
1106: If $X$ is a Euclidean building modeled on $(A,W)$, for each point
1107: $x\in X$ the space of directions $\Si_x(X)$ has {\em two structures}
1108: of a spherical building:
1109: 
1110: 
1111: 1. The {\em restricted building structure} which is modeled on the
1112: Coxeter complex $(S, W_{x})$, where $S=S_x(A)$ is the unit tangent
1113: sphere at $x$ and $W_{x}$ is the stabilizer of $x$ in the Coxeter
1114: group $W$. This building structure is thick.
1115: 
1116: 
1117: 2. The {\em unrestricted building structure} which is modeled on the
1118: Coxeter complex $(S, W_{sph})$, where $S=S_x(A)$ is the unit tangent
1119: sphere at $x$ and $W_{sph}$ is the linear part of the affine Coxeter
1120: group $W_{aff}$. This building structure is not thick, unless $x$ is
1121: a special vertex.
1122: 
1123: 
1124: Let $B$ be a spherical building modeled on a spherical Coxeter
1125: complex $(S, W_{sph})$. We say that two points $x, y\in B$ are
1126: antipodal, if $d(x,y)=\pi$; equivalently, they are antipodal points
1127: in an apartment $S'\subset B$ containing both $x$ and $y$. The
1128: quotient map $S\ra S/W_{sph}\cong\De_{sph}$ induces a canonical
1129: projection $\theta:B\ra \De_{sph}$ folding the building onto its
1130: model Weyl chamber. The $\theta$-image of a point in $B$ is called
1131: its {\em type}.
1132: 
1133: 
1134: \begin{rem}
1135: To define $\theta(x)$ pick an apartment $S'$ containing $x$ and a
1136: chart $\phi: S\to S'$. Then $\theta(x)$ is the projection of
1137: $\phi^{-1}(x)$ to  $S/W_{sph}\cong\De_{sph}$. We note that this is
1138: clearly independent of $S'$ and $\phi$.
1139: \end{rem}
1140: 
1141: 
1142: \begin{lem}
1143: \label{thetaproperties}
1144:  1. If $h: A\to A'$ is an isomorphism of apartments in $B$ (i.e.
1145: ${\phi'}^{-1}\circ h\circ \phi\in W$) then $\theta\circ h=\theta$.
1146: 
1147: 
1148: 2. If $x, x'\in B$ which belong to apartments $A, A'$ respectively
1149: and $-x\in A, -x'\in A'$ are antipodal to $x, x'$, then
1150: $\theta(x)=\theta(x')$ implies $\theta(-x)=\theta(-x')$.
1151: \end{lem}
1152: \proof (1) is obvious, so we prove (2). Pick an isomorphism $h: A\to
1153: A'$. Then (since $\theta(x)=\theta(x')$) there exists $w\in W\acts
1154: A'$ such that $w(h(x))=x'$. Hence $w\circ h(-x)=-w\circ(x)=-x'$. The
1155: claim now follows from 1. \qed
1156: 
1157: 
1158: 
1159: 
1160: 
1161: 
1162: \medskip
1163: We will regard {\em $n$-gons} $P$ in a building $X$ as maps $\nu:
1164: \{1,...,n\}\to X$, $\nu(i)=x_i$, where $x_i$ will be the vertices of
1165: $P$.  If $rank(X)\ge 1$ we can connect the consecutive vertices
1166: $x_i, x_{i+1}$ by shortest geodesic segments $\ol{x_i
1167: x_{i+1}}\subset X$ thus creating a {\em geodesic polygon} $[x_1,
1168: x_2,\ldots, x_n]$ in $X$ with the edges $\ol{x_i x_{i+1}}$. Observe
1169: that in case $x_i, x_{i+1}$ are antipodal, the edge $\ol{x_i
1170: x_{i+1}}$ is not unique.
1171: 
1172: 
1173: 
1174: 
1175: We say that two subsets $F, F'$ in a building $X$ are {\em
1176: congruent} if there exist apartments $A, A'$ in $X$ containing $F,
1177: F'$ resp., and an isomorphism $A\to A'$ of Coxeter complexes which
1178: carries $F$ to $F'$.
1179: 
1180: 
1181: \begin{conv}
1182: Suppose that $X$ is a spherical building. We will be considering
1183: only those geodesic triangles $T$ in $X$ for which the length of
1184: each geodesic side of $T$ is $\le \pi$.
1185: \end{conv}
1186: 
1187: 
1188: 
1189: 
1190: 
1191: \medskip
1192: Let $X, Y$ be buildings and $f: X\to Y$ a continuous map satisfying
1193: the following: For each alcove (in Euclidean case) or spherical
1194: chamber (in the spherical case) $a\subset X$, the image $f(a)$ is
1195: contained in an apartment of $Y$ and the restriction $f|a$ is
1196: an isometry or similarity. Then we call $f$ {\em differentiable} and
1197: define the {\em derivative} $df$ of $f$ as follows. Given a point
1198: $x\in X$ and $y=f(x)$, the derivative $df_x$ is a map $\Si_x(X)\to
1199: \Si_y(Y)$. For each $\xi\in \Si_x(X)$ let $\ol{xz}\subset X$ be a
1200: geodesic segment whose interior consists of regular points only and
1201: so that $\xi$ is the unit tangent vector to  $\ol{xz}$. Then $f$
1202: sends $\ol{xz}$ to a nondegenerate geodesic segment $\ol{y f(z)}$
1203: contained in an apartment $A\subset Y$. Then we let $df_x(\xi)\in
1204: \Si_y(Y)$ be the unit tangent vector to $\ol{y f(z)}$.
1205: 
1206: We will be also using the above definition in the setting when a building $Y$ is a Euclidean
1207: Coxeter complex $(A,W)$ and $h\in Dil(A,W)$. Then after letting $X:= h^*(A,W)$ we get an isometry $h: X\to Y$.
1208: 
1209: 
1210: \begin{conv}
1211: Throughout the paper we will be mostly using roman letters $x, y,
1212: z$, etc., to denote points in Euclidean buildings and Greek
1213: letters $\xi, \eta, \zeta$, etc., to denote points in spherical
1214: buildings. Sometimes however (e.g., in section \ref{converting})
1215: we will be working simultaneously with a spherical building $X$
1216: and its links $\Si_x(X)$, which are also spherical buildings. In
1217: this case we will use roman letters for points in $X$ and Greek
1218: letters for points  in $\Si_x(X)$.
1219: \end{conv}
1220: 
1221: \subsection{Generalized distances and lengths in buildings}
1222: \label{distances}
1223: 
1224: Let $(A,W)$ be a spherical or Euclidean Coxeter complex. The
1225: complete invariant of a pair of points $(x,y)\in A^2$ with respect
1226: to the action $W\acts A$, is its image $d_{ref}(x,y)$ under the
1227: canonical projection to $A\times A/W$. Following \cite{KLM2} we call $d_{ref}(x,y)$ the
1228: {\em refined distance from $x$ to $y$}. This notion carries over to
1229: buildings modeled on the Coxeter complex $(A,W)$: For a pair of
1230: points $(x,y)$ pick an apartment $A'$ containing $x,y$ and, after
1231: identifying $A'$ with the model apartment $A$, let $d_{ref}(x,y)$ be
1232: the projection of this pair to $A\times A/W$.
1233: 
1234: 
1235: 
1236: 
1237: If points $\xi, \eta$ in a spherical building are antipodal we will
1238: use $\pi$ for the refined distance $d_{ref}(\xi,\eta)$: This does
1239: not create much ambiguity since given apartment contains unique
1240: point antipodal to $\xi$.
1241: 
1242: 
1243: 
1244: 
1245: 
1246: 
1247: In the case of Euclidean Coxeter complexes there is an extra
1248: structure associated with the concept of refined length. Given a
1249: Euclidean Coxeter complex $(A, W_{aff})$, pick a special vertex
1250: $o\in A$. Then we can regard $A$ as a vector space $V$, with the
1251: origin $0=o$. Let $\De\subset A$ denote a Weyl chamber
1252:  of $W_{sph}$, the tip of $\De$ is at $o$.
1253: 
1254: 
1255: Then following \cite{KLM2}, we define the {\em $\De$-distance} between points of $(A,
1256: W_{aff})$ by composing $d_{ref}$ with the natural forgetful map
1257: \[ A\times A/W_{aff}\ra A/W_{sph}\cong\De .\]
1258: To compute the $\De$-distance $d_\De (x,y)$ we regard the oriented
1259: geodesic segment $\ol{xy}$ as a vector in $V$ and project it to
1260: $\De$. Again, the concept of $\De$-distance, carries over to the
1261: buildings modeled on $(A,W_{aff})$.
1262: 
1263: 
1264: \begin{defn}
1265: \label{maindef} Let $X$ be a thick Euclidean building. Define the
1266: set $D_n(X)\subset \De^n$ of $\De$-side lengths which occur for
1267: geodesic $n$-gons in $X$.
1268: \end{defn}
1269: 
1270: 
1271: It is one of the results of \cite{KLM2} that $D_n:=D_n(X)$ is a convex
1272: homogeneous polyhedral cone in $\De^n$, which depends only on $(A,W_{sph})$.
1273: The polyhedron $D_3$ in Theorem \ref{reform} is the polyhedron $D_3(X)$.
1274: The set of {\em stability} inequalities defining $D_n$
1275: is determined in \cite{KLM1} and
1276: \cite{BS}.
1277: 
1278: 
1279: 
1280: 
1281: \begin{thm}
1282: \label{klm2}\cite[Corollary 8.4]{KLM3}.
1283:  Let $X$ be a thick
1284: Euclidean building modeled on $(A,W_{aff})$. Suppose that $\al, \be,
1285: \ga\in P(R^\vee)$, $\al+\be+\ga\in Q(R^\vee)$ and $(\al,\be,\ga)\in
1286: D_3(X)$. Then there exists a geodesic triangle $T\subset X$ whose
1287: vertices are vertices of $X$ and the $\De$-side lengths are $\al,
1288: \be, \ga$.
1289: \end{thm}
1290: 
1291: 
1292: 
1293: \medskip
1294: Suppose that $p$ is a piecewise-linear path in a Euclidean building $X$. We say
1295: that $p$ is a {\em billiard path} if for each $t, s\in [0,1]$ the
1296: tangent vectors $p'(t), p'(s)$ have the same projection to the
1297: chamber $\De$ in the model apartment. If $p$ is a path which is the
1298: composition
1299: $$
1300: \ol{x_0 x_1}\cup ...\cup \ol{x_{m-1} x_{m}}
1301: $$
1302: of geodesic paths, then the $\De$-length of $p$ is defined as
1303: $$
1304: \Length(p):=\sum_{i=1}^m d_\De (x_{i-1}, x_i)
1305: $$
1306: where $d_\De (x,y)$ is the $\De$-distance from $x$ to $y$
1307: 
1308: 
1309: Each piecewise-linear path $p$ admits a unique representation
1310: $$
1311: p=p_1\cup...\cup p_n
1312: $$
1313: as a composition of maximal billiard subpaths so that
1314: $$
1315: \la_i= \Length (p_i).
1316: $$
1317: We define
1318: $$
1319: \length(p):= \ul{\la}= (\la_1,...,\la_n).
1320: $$
1321: Clearly, $\Length(p)$ is the sum of the vector components of
1322: $\ul{\la}$.
1323: 
1324: 
1325: 
1326: \subsection{The Hecke ring}
1327: \label{heck}
1328: 
1329: In this section we review briefly the definition of spherical
1330: Hecke rings and their relation to the geometry of Euclidean buildings; see
1331: \cite{Gross, KLM3} for more details.
1332: 
1333: Let $\K$ denote a locally compact field with discrete valuation
1334: $v$, and (necessarily) finite residue field of the order $q$. Let  $\mathcal{O}$
1335: be the subring of $\K$ consisting of elements with nonnegative
1336: valuation. Choose a uniformizer $\pi\in \mathcal{O}$.
1337: 
1338: Consider a connected reductive algebraic group $\ul{G}$ over $\K$.
1339: We fix a maximal split torus $\ul{T}\subset \ul{G}$ defined over
1340: $\mathcal{O}$. We put $G := \ul{G}(\K), K:= \ul{G}(\mathcal{O})$ and
1341: $T :=\ul{T}(\K)$. We let $\ul{B} \subset \underline{G}$ be a Borel subgroup
1342: normalized by $\underline{T}$ and set $B:= \underline{B}(\K)$.
1343: 
1344: Let $X$ denote the Bruhat-Tits building
1345: associated with the group $\ul{G}$; $o\in X$ is a distinguished special vertex
1346: stabilized by the compact subgroup $K$.
1347: 
1348: We also have  free abelian cocharacter group
1349: $X_*(\underline{T})$  of $\ul{T}$ whose rank equals $\dim(\underline{T})$.
1350: This group contains the set of coroots $R^{\vee}$ of the group $\ul{G}$.
1351: The roots are the characters of $\underline{T}$
1352: that occur in the adjoint representation on the Lie algebra of $\underline{G}$.
1353: The subset $R^+$ of the roots that occur in representation on the Lie algebra
1354: of $\underline{B}$ forms a positive system and the indecomposable elements
1355: of that positive system form a system of simple roots $\Phi$. We let $W$ denote
1356: the
1357: corresponding (finite) Weyl group.
1358: 
1359: The set of positive roots $\Phi$ determines a positive Weyl chamber $\Pi^+$ in
1360: $X_*(\underline{T})$, by
1361: $$
1362: \Pi^+ = \{\lambda \in X_*(\underline{T}): \ \<\lambda,\alpha\> \geq 0, \alpha
1363: \in \Phi\}.
1364: $$
1365: This chamber is a fundamental domain for the action of $W$ on
1366: $X_*(\underline{T})$.
1367: 
1368: We define a partial ordering on $\Pi^+$ by $\la \succ \mu$ iff the difference
1369: $\la-\mu$
1370: is a sum of positive coroots.
1371: 
1372: \begin{dfn}
1373: The (spherical) Hecke ring $\mathcal{H} = \mathcal{H}_{G}$ is
1374: the ring of all locally constant, compactly supported functions $f: G\lra \Z$
1375: which are $K$--biinvariant. The multiplication in $\mathcal{H}$ is by the
1376: convolution
1377: $$
1378: f \star g (z) = \int_G f(x) \cdot g(x^{-1}z)dx$$
1379: where $dx$ is the Haar measure on $G$ giving $K$ volume $1$.
1380: \end{dfn}
1381: 
1382: The ring $\mathcal{H}$ is commutative and associative (see e.g. \cite{Gross}).
1383: For $\lambda \in X_*(\ul{T})$ let
1384: $c_{\lambda}$ be the characteristic function of the corresponding
1385: $K$-double coset $\la(\pi)\in K\backslash G/K$.
1386: Then the functions $c_\la, \la\in \De$ freely generate ${\mathcal H}$ as an
1387: (additive) abelian group. Deep result of Satake \cite{Satake} relates the Hecke ring of $G$ and
1388: the representation ring of $G^\vee$.
1389: 
1390: 
1391: 
1392: The structure constants  for ${\mathcal H}$ are defined by the formula
1393: 
1394: \begin{equation}
1395: \label{heckesum}
1396: c_{\lambda} \star c_{\mu} = \sum_\nu
1397: m_{\lambda,\mu}(\nu) c_{\nu} =
1398: c_{\lambda + \mu} + \sum
1399: m_{\lambda,\mu}(\nu) c_{\nu},
1400: \end{equation}
1401: where the last sum is taken over all $\nu\in \Pi^+$ such that $\lambda + \mu
1402: \succ \nu$ and therefore is
1403: finite.
1404: 
1405: It turns out that the structure constants $m_{\lambda,\mu}(\nu)$
1406: are nonnegative integers which are polynomials in $q$ with integer
1407: coefficients. These constants are determined by the geometry of
1408: the building as follows. Given $\al, \be, \ga$ let
1409: $\mathcal{T}=\mathcal{T}_{\al,\be}(\ga)$ denote the (finite) set
1410: of geodesic triangles $[o, x, y]$ in the building $X$ which have
1411: the $\De$-side lengths $\al, \be, \ga^*$, so that $y$ is the
1412: projection of the point
1413: $$
1414: \gamma(\pi)\in T\subset G
1415: $$
1416: into $X$ under the map $g\mapsto g\cdot o$.
1417: Recall that $\ga$ as a cocharacter and therefore it defines a homomorphism
1418: ${\mathbb K}^*\to T$.
1419: 
1420: 
1421: \begin{thm}
1422: \cite[Theorem 9.11]{KLM3}.
1423: \label{trcount}
1424: $m_{\alpha,\beta}(\ga)$ equals the cardinality of $\mathcal{T}$.
1425: \end{thm}
1426: 
1427: \begin{rem}
1428: Instead of relating geometry of locally compact buildings to representation theory of $G^\vee$
1429: as it is done in this paper, one can use the non-locally compact building
1430: associated with the group $\ul{G}(\C((t)))$,
1431: as it is done, for instance, in \cite{GL}. It was shown in \cite{KLM2} that the choice of
1432: a field with discrete valuation (or, even, the entire Euclidean building)
1433: is irrelevant as far as the existence of triangles with the
1434: given $\De$-side lengths is concerned: What is important is the affine Weyl group.
1435: For our purposes, moreover, it is more convenient to work with local fields and locally-compact buildings.
1436: In particular, it allows us to compare structure constants of Hecke and representation rings.
1437: \end{rem}
1438: 
1439: 
1440: \section{Chains}
1441: \label{sectionchains}
1442: 
1443: 
1444: \subsection{Absolute chains}
1445: \label{absolute}
1446: 
1447: Let $R$ be a root system on a Euclidean vector space $V$,
1448: $W=W_{sph}$ be the finite Coxeter group associated with $R$, let
1449: $W_{aff}$ denote the affine Coxeter group associated to $R$. Our
1450: root system $R$ is actually the {\em coroot system} for the one
1451: considered by Littelmann in \cite{Littelmann2}. Accordingly, we will
1452: switch weights to coweights, etc. We pick a Weyl chamber $\Del$ for
1453: $W$, this determines the positive roots and the simple roots in $R$.
1454: Let $-\Del$ denote the negative chamber.
1455: 
1456: 
1457: We get the Euclidean Coxeter complex $(A,W_{aff})$, where $A=V$ and
1458: the spherical Coxeter complex $(S,W)$ where $S$ is the
1459: unit sphere in $V$. By abusing notation we will also use the
1460: notation $\De, -\De$ for the positive and negative chambers in
1461: $(S,W)$. We will use the notation $(A,W,-\De)$ for a
1462: Euclidean/spherical Coxeter complex with chosen negative chamber.
1463: More generally, we will use the notation $(A,W_{aff},a)$ for a Euclidean
1464: Coxeter complex with chosen alcove $a$.
1465: 
1466: 
1467: \begin{figure}[tbh]
1468: \centerline{\epsfxsize=4.5in \epsfbox{Fc.eps}} \caption{\sl A
1469: chain.} \label{Fc}
1470: \end{figure}
1471: 
1472: 
1473: \begin{defn}\label{D0}
1474:  A $W$--{\em chain} in
1475: $(A,W,-\De)$ is a finite sequence $(\eta_0,...,\eta_m)$ of nonzero
1476: vectors in $V$ so that for each $i=1,...,m$ there exists a positive
1477: root $\be_i\in R^+$ so that the corresponding reflection
1478: $\tau_i:=\tau_{\be_i}\in W$ satisfies
1479: 
1480: 
1481: 1. $\tau_i(\eta_{i-1})= \eta_{i}$.
1482: 
1483: 
1484: 2. $\be_i(\eta_{i-1}) <0$.
1485: 
1486: 
1487: \noindent Sometimes we will refer to a chain as a $(V,W,-\De)$-chain to
1488: emphasize the choice of $V$, $W$ and $\De$. When the choice of $W$
1489: is clear we will frequently refer to $W$-chains as {\em chains}.
1490: \end{defn}
1491: 
1492: \begin{rem}
1493: One could call these chains $(V,W,\De)$-chains instead,
1494: but this definition would not generalize to affine chains.
1495: \end{rem}
1496: 
1497: 
1498: 
1499: 
1500: Recall that $\tau_i$ is the reflection in the wall
1501: $H_i=\{\be_i=0\}$. More geometrically one can interpret the
1502: condition (2) by saying that the wall $H_i$ separates the negative
1503: chamber $-\De$ from the vector $\eta_i$. In other words, the
1504: reflection $\tau_i$ moves the vector $\eta_{i-1}$ ``closer'' to the
1505: positive chamber.
1506: 
1507: \bigskip
1508: The concept of a chain generalizes naturally to Euclidean Coxeter complexes $(A,W_{aff})$. Pick an
1509: alcove $a$ in $A$.
1510: 
1511: \begin{defn}
1512: \label{DA}
1513: An affine {\em chain} in $(A,W_{aff},a)$ is a finite sequence $(\eta_0,...,\eta_m)$ of
1514: elements in $A$ so that for each $i=1,...,m$ there exists a reflection $\tau_i\in W_{aff}$ such that
1515: 
1516: 1. $\tau_i(\eta_{i-1})= \eta_{i}$.
1517: 
1518: 
1519: 2. The hyperplane $H_i\subset A$ fixed by $\tau_i$ separates $a$ from $\eta_i$.
1520: \end{defn}
1521: 
1522: 
1523: 
1524: We now return to the chains as in definition \ref{D0}.
1525: By restricting vectors $\eta_i$ to have unit length we define chains
1526: in the spherical Coxeter complex, see Figure \ref{Fc}.
1527: 
1528: 
1529: \begin{defn}
1530: The points $\eta_i$ as in Definition \ref{D0} will be called {\em vertices} of the
1531: chain. We will say that the chain begins at $\eta_0$ and {\em ends}
1532: at $\eta_m$, or that this chain is {\em from} $\eta_0$ {\em to}
1533: $\eta_m$. We refer to a subsequence $(\eta_i,
1534: \eta_{i+1},...,\eta_m)$ as a {\em tail} of the chain.
1535: \end{defn}
1536: 
1537: 
1538: We will refer to the number $m$ as the {\em length} of the chain. A
1539: chain $(\eta_i)$ is called {\em simple} if it has length 1. Set
1540: $$
1541: dist_W(\nu,\mu)=dist(\nu, \mu)$$ to be the maximal length $m$ of a
1542: $W$-chain which begins at $\nu$ and ends at $\mu$.
1543: 
1544: 
1545: 
1546: 
1547: 
1548: Given a chain $(\eta_0,...,\eta_m)$ we define a {\em subdivision} of
1549: this chain to be a new chain in $(A,W)$ which is still a chain
1550: from $\eta_0$ to $\eta_m$ and which contains all the vertices of
1551: the original chain.
1552: 
1553: 
1554: The concept of chain determines a partial order on the $W$-orbits in
1555: $V$:
1556: 
1557: 
1558: \begin{defn}
1559: \cite[page 509]{Littelmann2}
1560: \label{order}
1561:  For a pair of nonzero vectors $\nu, \mu\in V$ which
1562: belong to the same $W$-orbit, write $\nu\ge_{W} \mu$ (or simply
1563: $\nu\ge \mu$) if there exists a $W$-chain from $\nu$ to $\mu$.
1564: Accordingly, $\nu >\la$ if $\nu\ge \la$ and $\nu \ne \la$.
1565: \end{defn}
1566: 
1567: 
1568: 
1569: 
1570: \begin{lem}
1571: \label{easy} Suppose that $\nu \ge \mu$ and $\al$ is a positive root
1572: such that $\al(\mu)\le 0$. Then
1573: $$
1574: \nu\ge \tau_\al(\mu).
1575: $$
1576: \end{lem}
1577: \proof If $\al(\mu)=0$ then $\tau_\al(\mu)=\mu$ and there is nothing
1578: to prove. Thus we assume that $\al(\mu)<0$. Consider a chain
1579: $(\nu=\nu_0,...,\nu_s=\mu)$. Then, since $\al(\mu)<0$, we also have
1580: $\al(\tau_\al(\mu))>0$ and thus we get a longer chain
1581: $$
1582: (\nu=\nu_0,...,\nu_s=\mu, \tau(\mu)). \qed
1583: $$
1584: 
1585: 
1586: The word metric $d_W$ on the finite Coxeter group $W$ defines the
1587: {\em length function}
1588: $$
1589: \ell: W\cdot \la \to \N
1590: $$
1591: by
1592: $$
1593: \ell(\mu):= \min\{ d_W(w,1):  w\in W, w^{-1}(\mu)\in \De\}.
1594: $$
1595: 
1596: 
1597: \begin{prop}
1598: \label{wordlength} If $\nu > \mu$ then $\ell(\nu)> \ell(\mu)$.
1599: \end{prop}
1600: \proof It suffices to prove the assertion in the case when
1601: $$
1602: \mu=\tau(\nu), \tau=\tau_\be,
1603: $$
1604: where $\be$ is a positive root, $\be(\nu)<0, \be(\mu)>0$. If $W\cong
1605: \Z/2$ the assertion is clear, so we suppose that it is not the case.
1606: Then we can embed the Cayley graph $\Ga$ of $W$ as a dual graph to
1607: the tessellation of $V$ by the Weyl chambers of $W$. Suppose that
1608: $\nu \in w^{-1}(\De)$, then the wall $H_{\be}=\{\be=0\}$ separates
1609: $w^{-1}(\De)$ from $\De$, where $w\in W$ is the shortest element
1610:  such that $w(\nu)\in \De$. Let $p: [0,1]\to \Ga$
1611: denote the shortest geodesic from $1$ to $w$ in $\Ga$. The path $p$
1612: crosses the wall $H$ at a point $x=p(T)$. We construct a new path
1613: $q$  by
1614: $$
1615: q|[0,T]=p|[0,T], \quad q|[T, 1]= w\circ p|[T,1].
1616: $$
1617: The path $q$ connects $1\in \De$ to the Weyl chamber $\tau w(\De)$
1618: containing $\mu$. This path has a break-point at $x$, which is not a
1619: vertex of the Cayley graph. Therefore, by eliminating the
1620: backtracking of $q$ at $x$, we obtain a new path which connects $1$
1621: to $\tau w(\De)$ and whose length is one less than the length of
1622: $p$. \qed
1623: 
1624: 
1625: \begin{cor}
1626: \label{maxi} The length of a chain in $V$ does not exceed the
1627: diameter of the Cayley graph of $W$.
1628: \end{cor}
1629: 
1630: 
1631: \begin{cor}
1632: \label{end} Suppose that $\nu\in \De$ and $\nu\ge \mu$. Then
1633: $\mu=\nu$.
1634: \end{cor}
1635: \proof Since $\nu\in \De$, $\ell(\nu)=0$. Hence by Proposition \ref{wordlength},
1636: $\ell(\mu)= 0$, which implies  that $\mu=\nu$. \qed
1637: 
1638: 
1639: 
1640: \begin{lem}
1641: \label{2orders}
1642: Suppose that $\nu=w(\mu)\ne 0$ for some $w\in W$ and $\mu\rhd \nu$. Then
1643: $\nu \ge \mu$.
1644: \end{lem}
1645: \proof Let $\De$ be the positive Weyl chamber.
1646: Suppose that $\De_0$ is a chamber containing $\nu$ and $\De'$
1647: is a chamber containing $\mu$. These chambers are non-unique, but
1648: we can choose them in such a way that for all $\xi\in \De_0, \xi'\in \De'$
1649: $$
1650: \xi'\rhd \xi.
1651: $$
1652: In other words, if a wall separates $\De'$ from $\De$ then it also separates $\De_0$ from $\De$.
1653: Let
1654: $$
1655: (\De_0, \De_1,...,\De_m=\De')
1656: $$
1657: be a gallery of Weyl chambers, i.e. for each $i$, $\De_i\cap
1658: \De_{i+1}$ is a codimension 1 face $F_i$ of $\De_i, \De_{i+1}$. We choose this gallery to have
1659: the shortest length, i.e. so that the number $m$ is minimal.
1660: Let $H_i$ be the wall containing $F_i$ and $\tau_i$ be the reflection in $H_i$.
1661: We claim that the sequence
1662: $$
1663: \nu=\eta_0, \eta_1=\tau_1(\eta_0),..., \mu=\eta_m=
1664: \tau_m(\eta_{m-1}),
1665: $$
1666: after deletion of equal members, is a chain.
1667: 
1668: Our proof is by induction on $m$. If $m=0$ and $\nu=\mu$, there is
1669: nothing to prove. Suppose the assertion holds for $m-1$, let us
1670: prove it for $m$. We claim that for all points $\xi_1\in \De_1,
1671: \xi_0\in \De_0$, $\xi_1 \rhd \xi_0$. Indeed, otherwise the wall
1672: $H_1$ does not separate $\De_0$ from $\De$, but separates $\De_1$
1673: from $\De$. Then $H_1$ does not separate $\De_m$ from $\De$
1674: either. Thus, as in the proof of Proposition \ref{wordlength}, we can
1675: replace the gallery $(\De_0, \De_1,...,\De_m)$ with a shorter
1676: gallery connecting $\De_0$ to $\De'$, contradicting minimality of
1677: $m$. Now, clearly,
1678: $$
1679: \eta_0\ge \eta_1, \eta_0\rhd \eta_1.
1680: $$
1681: Therefore, by the induction
1682: $$
1683: \eta_0\ge \eta_1 \ge \eta_m \Rightarrow \nu=\eta_0\ge \eta_m=\mu.  \qed
1684: $$
1685: 
1686: \begin{rem}
1687: The converse to the above lemma is false for instance for
1688: the root system $A_2$. See Figure \ref{Fchain}, where $\eta_0\ge \eta_1$ but  $\eta_0\ntriangleright \eta_1$.
1689: \end{rem}
1690: 
1691: 
1692: 
1693: 
1694: \begin{figure}[tbh]
1695: \centerline{\epsfxsize=4.5in \epsfbox{ch.eps}} \caption{\sl A
1696: chain.} \label{Fchain}
1697: \end{figure}
1698: 
1699: 
1700: As a corollary of Lemma \ref{2orders} we obtain:
1701: 
1702: \begin{lem}
1703: \label{constructingachain}
1704: Let $\nu\in V\setminus \{0\}$ and let $\mu$ be the unique vector in
1705: $W\cdot \nu$ which belongs to $\De$. Then $\nu\ge \mu$.
1706: \end{lem}
1707: \proof Clearly, $\mu \rhd \nu$. Then the assertion follows from
1708: Lemma \ref{2orders}.\qed
1709: 
1710: 
1711: 
1712: 
1713: 
1714: 
1715: 
1716: \begin{defn}
1717: [Maximality condition] \label{maximality}
1718:   We say that a chain
1719: $(\eta_0, \eta_1,..., \eta_m)$ is  {\em maximal} if it cannot be
1720: {\em subdivided} into a longer $W$-chain. Equivalently, \newline
1721: $dist(\eta_i,\eta_{i+1})=1$ for each $i=0,...,m-1$.
1722: \end{defn}
1723: 
1724: 
1725: 
1726: 
1727: 
1728: 
1729: 
1730: 
1731: 
1732: 
1733: \begin{lem}
1734: \label{simplemax}
1735: Suppose that $\nu\ge \mu$, and there exists a simple root $\be$ such that
1736: $\tau_\be(\nu)=\mu$ and  $\be(\nu)<0$. Then $dist(\nu,\mu)=1$.
1737: \end{lem}
1738: \proof Consider a chain from $\nu$ to $\mu$, i.e. a sequence of vectors $\nu=\nu_0, \nu_1,...,
1739: \nu_s=\mu$ and positive roots $\be_1,...,\be_s$ so that
1740: $$
1741: \nu_i= \tau_{\be_i}(\nu_{i-1}) \quad \hbox{and}\quad
1742: \be_i(\nu_{i-1}) <0, i=1,...,s.
1743: $$
1744: Then
1745: $$
1746: \mu-\nu= -\be(\nu) \be^\vee,
1747: $$
1748: and
1749: $$
1750: \mu-\nu= \sum_{i=1}^s -\be_{i-1} (\nu_i)\be^\vee_i.
1751: $$
1752: Thus
1753: $$
1754: \be= \sum_{i=1}^s \frac{ \< \nu_i, \be_i^\vee\>}{\<\nu, \be^\vee\>}
1755: \be_{i-1}
1756: $$
1757: i.e. the simple root $\be$ is a positive linear combination of
1758: positive roots. It follows that $s=1$ and $\be_1=\be$. \qed
1759: 
1760: 
1761: 
1762: 
1763: 
1764: 
1765: 
1766: 
1767: 
1768: 
1769: 
1770: \subsection{Relative chains}
1771: 
1772: 
1773: {\bf 1. Chains relative to a root subsystem.}
1774: 
1775: 
1776: \medskip
1777: Let $R$ be a root system on $V$ with the set of simple roots $\Phi$,
1778: $W$ be the corresponding Weyl group. Let $\Phi'\subset \Phi$ be a
1779: subset, $W'\subset W$ the corresponding reflection subgroup and
1780: $\De'$ the positive chamber for $W'$, defined by the property that
1781: all simple roots $\al\in \Phi'$ are nonnegative on $\De'$. Thus we
1782: will be (frequently) considering $(V,W',-\De')$-chains rather than $(V,W,-\De)$-chains.
1783: In this paper we will be using subgroups $W'$ which are stabilizers
1784: of points $x$ in $(A,W_{aff})$ (where $W=W_{sph}$). Then we will
1785: think of a relative $(V, W')$-chain as a chain in the tangent space
1786: $T_x A$ of the point $x$.
1787: 
1788: 
1789: \begin{lem}
1790: \label{simplesubsystem} Given a nonzero vector $\eta\in V$ there
1791: exists a $W'$-chain \newline $(\eta=\eta_0,...,\eta_m)$ with
1792: $\eta_m\in \De'$, so that the chain $(\eta_i)$ is maximal with
1793: respect to the original root system $R$.
1794: \end{lem}
1795: \proof The proof is by induction on the number $r=r(\eta)$ of simple
1796: roots in $\Phi'$ which are negative on the vector $\eta$. If
1797: $r(\eta)=0$ then $\eta\in \De'$ and there is nothing to prove.
1798: Suppose the assertion holds for all $\eta$ with $r(\eta)\le k$. Let
1799: $\eta$ be such that $r(\eta)=k+1$. Pick a root $\be\in \Phi'$ such
1800: that $\be(\eta)<0$. Then for the vector
1801: $$
1802: \zeta:= \tau_{\be}(\eta)
1803: $$
1804: we have:
1805: $$
1806: \zeta= \eta- \be(\eta)\be^\vee.
1807: $$
1808: Clearly, $\beta(\zeta) >0$. Thus the pair $(\eta,\zeta)$ is a
1809: $W'$-chain; this chain is maximal as a $W$-chain by Lemma
1810: \ref{simplemax}, since it is defined using a simple reflection in
1811: $W$.
1812: 
1813: 
1814: For each simple root $\al\in \Phi'\setminus \{\be\}$ which is
1815: nonnegative on $\eta$
1816: $$
1817: \al(\zeta) = \al(\eta)- \be(\eta) \al(\be^\vee)\ge \al(\eta) \ge 0.
1818: $$
1819: Therefore $r(\zeta)< r(\eta)$ and we are done by the induction. \qed
1820: 
1821: 
1822: 
1823: 
1824: \begin{lem}
1825: \label{ifff} Suppose that $W'\subset W$ is a reflection subgroup as
1826: above. For any two vectors $\al, \del$ the following are equivalent:
1827: 
1828: 
1829: 1. There exists $\be, \ga$ so that $\al \ge_{W'} \be \sim_{W} \ga
1830: \ge_{W'}  \del$.
1831: 
1832: 
1833: 2. There exists $\be$ so that $\al \ge_{W'}  \be \sim_{W} \del$.
1834: 
1835: 
1836: 3. There exists $\ga$ so that $\al \sim_{W} \ga \ge_{W'}  \del$.
1837: 
1838: 
1839: \noindent Here $\sim$ is the relation from Definition \ref{relation}.
1840: \end{lem}
1841: \proof It is clear that 2 $\Rightarrow$ 1 and 3 $\Rightarrow$ 1. We
1842: will prove that 1 $\Rightarrow$ 2, since the remaining implication
1843: is similar. We have chains
1844: $$
1845: (\al=\eta_0,...,\eta_m=\be), \quad \eta_i=\tau_i(\eta_{i-1}), \ i=1,...,m,
1846: $$
1847: $$
1848: (\ga=\eta_0',...,\eta_s'=\del), \quad \eta_i'=\tau_i'(\eta_{i-1}'), \ i=1,...,s.
1849: $$
1850: Then we can extend the first chain to
1851: $$
1852: (\al=\eta_0,...,\eta_m=\be, \tau_1'(\be)=\eta_{m+1},...,
1853: \tau_s'(\eta_{m+s-1})=\eta_{m+s}=:\eps).
1854: $$
1855: After discarding equal members of this sequence we obtain a chain
1856: from $\al$ to $\eps$. Since $\be, \ga$ belong to the same chamber,
1857: the vectors $\eps$ and $\del=\tau_s'\circ ... \circ \tau_1'(\ga)$
1858: also belong to the same chamber. Therefore we obtain
1859: $$
1860: \al\ge \eps \sim \del. \qed
1861: $$
1862: 
1863: 
1864: \begin{defn}
1865: We will write $\al\gtrsim_{W'}  \del$ if one of the equivalent
1866: conditions in the above lemma holds. We will frequently omit the
1867: subscript $W'$ in this notation when the choice of the subgroup $W'$
1868: is clear or irrelevant.
1869: \end{defn}
1870: 
1871: 
1872: The reason for using the relation $\sim$ with respect to $W$ rather
1873: than its subgroup $W'$ (as may seem more natural) is that we will be
1874: taking limits under which the subgroup $W'$ is increasing (but the
1875: limit is still contained in $W$). Such limits clearly preserve the
1876: relation $\gtrsim_{W'}$, but not the relation defined using
1877: $\sim_{W'}$.
1878: 
1879: 
1880: 
1881: 
1882: The next corollary immediately follows from Lemma \ref{ifff}:
1883: 
1884: 
1885: \begin{cor}
1886: \label{iff}
1887:  $\al \gtrsim \del \iff -\del \gtrsim -\al$.
1888: \end{cor}
1889: 
1890: 
1891: \begin{lem}
1892: \label{hard} Suppose that $\nu \gtrsim \mu$ and $\al\in \Phi$ is
1893: such that $\al(\nu)<0$, $\al(\mu)\ge 0$. Then
1894: $$
1895: \tau_\al(\nu)\gtrsim \mu.
1896: $$
1897: \end{lem}
1898: \proof Let $\la$ be such that
1899: $$
1900: \nu \ge \la \sim \mu.
1901: $$
1902: Then $\al(\la)\ge 0$ and it follows that $\la\ne \nu$, i.e. $\nu >
1903: \la$. By applying \cite[Lemma 4.3]{Littelmann2} we get
1904: $$
1905: \tau_\al(\nu) \ge \la \sim \mu. \qed
1906: $$
1907: 
1908: 
1909: \medskip
1910: {\bf 2. Chains relative to positive real numbers ($a$-chains in the
1911: sense of Littelmann).}
1912: 
1913: Let $a$ be a positive real number
1914: and let $\nu, \mu\in V$ be nonzero vectors in the same $W$-orbit.
1915: 
1916: 
1917: \begin{defn}
1918: [P.~Littelmann, \cite{Littelmann2}] \label{achain}  {\em An
1919: $a$-chain} for $(\nu,\mu)$ is a chain $(\la_0,\ldots,\la_s)$ which
1920: starts at $\nu$, ends at $\mu$ and satisfies
1921: 
1922: 
1923: (i) For each $i>0$ we have
1924: $$
1925: t_i:=\be_i(a\la_{i-1})\in \Z,
1926: $$
1927: where $\la_i=\tau_{\be_i}(\la_{i-1})$ as in the Definition
1928: \ref{D0}.
1929: 
1930: 
1931: (ii) For each $i$, $dist(\la_{i-1}, \la_i)=1$.
1932: \end{defn}
1933: 
1934: 
1935: \begin{rem}
1936: Our root system $R$ is the coroot system for the one considered by
1937: Littelmann.
1938: \end{rem}
1939: 
1940: 
1941: Our goal is to give this definition a somewhat more geometric
1942: interpretation. In particular, we will see that the concept of an
1943: $a$-chain is a special case of the concept of a chain relative to a
1944: root subsystem.
1945: 
1946: 
1947: 
1948: 
1949: The root system $R$ defines an affine Coxeter complex $(A, W_{aff})$
1950: on $A$. Let $x\in P(R^\vee)$ be a special vertex; set $x_i:=
1951: x+a\la_i$, $\tau_i:=\tau_{\be_i}$, $i=0,...,s$. Thus $t_i=
1952: \be_i(x_{i-1})$. Note that $t_i\in \Z$ iff $\be_i(x_{i-1})\in \Z$.
1953: 
1954: 
1955: \begin{prop}
1956: $\be_i( x_0)\in \Z$ for each $i=1, 2,...,s$.
1957: \end{prop}
1958: \proof It suffices to consider the case $x=0$. We have:
1959: $$
1960: x_{i-1}= \tau_{i-1}(x_{i-2})= x_{i-2} - \be_{i-1}( x_{i-2})
1961: \be_{i-1}^\vee= x_{i-2} - t_{i-1} \be_{i-1}^\vee= \ldots = x_0-
1962: \sum_{j=1}^{i-1}  t_j \be_{j}^\vee.
1963: $$
1964: Hence
1965: $$
1966: t_i= \be_i( x_{i-1})= \be_i( x_0) -  \sum_{j=1}^{i-1}  t_j
1967: \be_i(\be_{j}^\vee).
1968: $$
1969: Since $t_i\in \Z$ and $\be_i(\be_{j}^\vee)\in \Z$ for all $j$, it
1970: follows that $\be_i( x_0)\in \Z$. \qed
1971: 
1972: 
1973: 
1974: 
1975: \medskip
1976: We define the integers $k_i:= \be_i(x_0)$ and the affine walls
1977: $$
1978: H_i:=H_{\be_{i},k_{i}}= \{ v\in V: \be_{i}(v)=k_i\}.
1979: $$
1980: The reflection $\si_i$ in  the wall $H_i$ belongs to the group
1981: $W_{aff}$, its linear part is the reflection $\tau_i\in W_{sph}$,
1982: $i=1,...,s$.
1983: 
1984: 
1985: 
1986: 
1987: The argument in the above proof can be easily reversed
1988: and hence we get
1989: 
1990: 
1991: \begin{cor}
1992: \label{C1} The integrality condition (i) is equivalent to the
1993: assumption that the point $x_0$ lies on the intersection of walls
1994: $H_i$ of the Euclidean Coxeter complex $(A,W_{aff})$, where each
1995: $H_i$ is parallel to the reflection hyperplane of $\tau_i$.
1996: Equivalently, the $W$-chain $(\la_0,...,\la_s)$ is actually a
1997: $W'$-chain, where $W'=W_{x_0}$ is the stabilizer of $x_0$ in
1998: $W_{aff}$.
1999: \end{cor}
2000: 
2001: 
2002: Therefore, identify the vectors $\la_i$ with vectors in the tangent
2003: space $V':= T_{x_0}(A)$, let $\De'\subset V'$ denote the Weyl
2004: chamber of $W'$ which contains the (parallel transport of the)
2005: positive chamber $\De$. We obtain
2006: 
2007: 
2008: \begin{prop}
2009: \label{eq}  Littelmann's definition of an $a$-chain is equivalent to
2010: the conjunction of
2011: 
2012: 
2013: 1. $(\la_0,...,\la_s)$ is a chain in $(V',W',-\De')$.
2014: 
2015: 
2016: 2. This chain is maximal as a $W$-chain.
2017: \end{prop}
2018: 
2019: 
2020: Thus the choice of the real number $a$ amounts to choosing a
2021:  Coxeter subcomplex $(V', W')$ in $(V,W_{sph})$. The reader will also note
2022: the discrepancy between (1) and (2): The chain condition refers to
2023: the restricted Coxeter complex $(V', W')$, while the maximality
2024: condition refers to the unrestricted one, $(V,W_{sph})$. This is the
2025: key difference between {\em LS paths} and {\em Hecke paths}.
2026: 
2027: 
2028: \begin{rem}
2029: \label{R1} Note that both conditions (1) and (2) are vacuous if
2030: $x_0$ is a special vertex in the Euclidean Coxeter complex
2031: (equivalently, if $a\nu$ is a coweight): If
2032: $$
2033: (\la_0, \la_1,...,\la_s)
2034: $$
2035: is a $W'$-chain, since $W'\cong W_{sph}$,  we can subdivide this
2036: chain to get a longer chain
2037: $$
2038: (\la_0=\la'_0, \la_1',...,\la_{m-1}', \la_m'=\la_s)
2039: $$
2040: between $\la_0$ and $\la_s$ which satisfies the unit distance
2041: condition  $dist(\la_i', \la_{i+1}')=1$ for all $i$.
2042: \end{rem}
2043: 
2044: 
2045: 
2046: 
2047: 
2048: 
2049: 
2050: 
2051: 
2052: 
2053: \subsection{Hecke paths}
2054: 
2055: 
2056: The goal of this section is to introduce a class of piecewise-linear paths which
2057: satisfy a condition similar to Littelmann's definition of LS (and
2058: generalized LS) paths. These paths ({\em Hecke paths}) play the role
2059: in the problem of computing structure constants for spherical Hecke
2060: algebras which is analogous to the role that LS paths play in the
2061: representation theory of complex semisimple Lie groups.
2062: 
2063: 
2064: \medskip
2065: Let $(A, W_{aff}, -\De)$ be a Euclidean Coxeter complex
2066: corresponding to a root system $R$, with fixed negative chamber
2067: $-\De$.  Let $p\in \tilde{\mathcal P}$ be a path equal to the
2068: composition
2069: $$
2070: \ol{x_{1} x_{2}}\cup ... \cup \ol{x_{n-1} x_{n}}.
2071: $$
2072: For each vertex $x=x_i, i=2,...,n-1$ we define the unit tangent
2073: vectors $\xi, \eta \in S_x$ to the segments $\ol{x_ix_{i-1}},
2074: \ol{x_ix_{i+1}}$.
2075: 
2076: 
2077: 
2078: 
2079: \begin{defn}
2080: \label{D3} We say that the path $p$ {\em satisfies the chain
2081: condition} if for each $x=x_i, i=2.,...,n-1$ there exists a unit
2082: vector $\mu$ so that
2083: 
2084: 
2085: 1.
2086: $$
2087: -\xi \ge_{W_x} \mu, \hbox{in the spherical Coxeter complex~}
2088: (S_x,W_x,\De).
2089: $$
2090: 2. $\mu \sim \eta$ in the (unrestricted) spherical Coxeter complex
2091: $(S, W_{sph})$, i.e. for each root $\al\in R$ we have
2092: $$
2093: \al(\mu) \ge 0 \iff \al(\eta) \ge 0.
2094: $$
2095: In other words, for each $t\in [0, 1]$ we have
2096: $$
2097: p'_-(t)\gtrsim_{W_{p(t)}} p'_+(t).
2098: $$
2099: \end{defn}
2100: 
2101: 
2102: \noindent Intuitively, at each break-point $p(t)$ the path $p$
2103: ``turns towards the positive chamber''. In what follows we will use
2104: the notation ${\mathcal P}_{chain}$ for the set of all paths $p\in
2105: {\mathcal P}$ satisfying the chain condition.
2106: 
2107: 
2108: 
2109: 
2110: \begin{defn}
2111: \label{Hecke} A path $p\in \t{\mathcal P}$ is called a {\em Hecke
2112: path} if it is a billiard path which satisfies the chain condition, i.e.
2113: for each $t$
2114: $$
2115: p'_-(t)\ge_{W_{p(t)}} p'_+(t).
2116: $$
2117: \end{defn}
2118: 
2119: 
2120: 
2121: 
2122: 
2123: 
2124: Below is example of a class of Hecke paths. Suppose that $p\in
2125: {\mathcal P}$ and for each $t\in [0, 1]$ either $p$ is smooth at $t$
2126: or there exists a reflection $\tau\in W_{p(t)}$ so that the
2127: derivative of $\tau$ equals $\tau_\be$, $\be \in R^+$ and
2128: 
2129: 
2130: 1. $d\tau_{p(t)} (p'_-(t))=p'_+(t)$.
2131: 
2132: 
2133: 
2134: 
2135: 2. $\be(p'_-(t))<0$, $\be(p'_+(t))>0$.
2136: 
2137: 
2138: Then $p$ is a Hecke path with the length of each chain in
2139: $(S_{p(t)}, W_{p(t)})$ equal $0$ or $1$. See Figure \ref{Fa}.
2140: 
2141: 
2142: \begin{defn}
2143: We say that $p$ satisfies the {\em simple chain} condition if at
2144: each break-point $x=p(t)$ the chain can be chosen to be simple, i.e.
2145: of length $1$.
2146: \end{defn}
2147: 
2148: 
2149: 
2150: 
2151: 
2152: 
2153: \begin{figure}[tbh]
2154: \centerline{\epsfxsize=3.5in \epsfbox{Fa.eps}} \caption{\sl A
2155: billiard path satisfying the simple chain condition.} \label{Fa}
2156: \end{figure}
2157: 
2158: 
2159: In what follows we will also need
2160: 
2161: \begin{defn}
2162: \label{gH}
2163: Suppose that $p$ is a path satisfying the chain condition. We call $p$ a {\em generalized Hecke path} if
2164: each geodesic segment in $p$ (regarded as a vector)
2165: is an integer multiple of some $w\varpi$ where $w\in W$ and $\varpi=\varpi_i$
2166: is one of the fundamental coweights.
2167: \end{defn}
2168: 
2169: 
2170: 
2171: 
2172: 
2173: \subsection{A compactness theorem}
2174: 
2175: 
2176: Pick $\eps>0$. We define the subset ${\mathcal P}_{m,\eps}\subset
2177: {\mathcal P}$ consisting of paths $p$ with
2178: $$
2179: \length(p)=\ul{\la}=(\la_1,...,\la_m)$$
2180:  so that for each $i$,
2181: \begin{equation}
2182: \label{bdd}
2183: \eps\le |\la_i|\le \eps^{-1}.
2184: \end{equation}
2185: 
2186: 
2187: 
2188: 
2189: \begin{thm}
2190: \label{compactness} For each $\eps>0$ the set ${\mathcal
2191: P}_{chain,m,\eps}:= {\mathcal P}_{chain}\cap {\mathcal P}_{m,\eps}$
2192: is compact in ${\mathcal P}_{m,\eps}$.
2193: \end{thm}
2194: \proof Suppose that $p\in {\mathcal P}_{chain}$ is a concatenation
2195: of $m$ billiard paths $p_i$. Then the number of breaks in the broken
2196: geodesic $p_i$ is bounded from above by a constant $c$ equal to the
2197: length of a maximal chain in the Bruhat order of the finite Weyl
2198: group $W_{sph}$, see Corollary \ref{maxi}. This immediately implies
2199: that the subset ${\mathcal P}_{chain,m,\eps}$ is precompact in
2200: ${\mathcal P}$. What has to be proven is that this subset is closed.
2201: 
2202: 
2203: \bigskip
2204: Let $\la_{ij}\in V$ be vectors of nonzero length so that
2205: $$
2206: p_i= \pi_{\la_{i1}} * \pi_{\la_{i2}}*...*\pi_{\la_{in}}
2207: $$
2208: are Hecke paths, $i=1, 2,...$. We suppose that
2209: $$
2210: \lim_{i\to\infty} \la_{ij}=0, j=2,...,n-1
2211: $$
2212: and
2213: $$
2214: \lim_{i\to\infty} \la_{i1}=\la_{\infty,1}, \lim_{i\to\infty} \la_{in}=\la_{\infty, n}
2215: $$
2216: are nonzero vectors. It is clear that
2217: $$
2218: \lim_i p_i= p_{\infty}:= \pi_{\la_{\infty,1}} * \pi_{\la_{\infty,n}}.
2219: $$
2220: 
2221: 
2222: 
2223: \begin{lem}
2224: \label{1}
2225: Under the above conditions the path $p_{\infty}$ is again a Hecke path.
2226: \end{lem}
2227: \proof Let $x:= \la_{\infty,1}$. We need to check that the unit vectors $\bar\la_{\infty,1}, \bar\la_{\infty,n}$ satisfy
2228: $$
2229: \bar\la_{\infty,1} \ge_{W_x} \bar\la_{\infty,n}.
2230: $$
2231: Here and below, $W_x$ is the stabilizer of $x\in V$ in the Coxeter group $W_{aff}$.
2232: 
2233: 
2234: Let $x_{ij}$ denote the break-point of $p_i$ which is the concatenation point between $\la_{i,j}$ and $\la_{i,j+1}$. Then
2235: $$
2236: \bar\la_{i,j} \ge_{W_{x_{ij}}} \bar\la_{i,j+1},
2237: $$
2238: $\lim_i x_{ij}=x$. If $\si_{ij}\in W_{aff}$ is a reflection fixing $x_{ij}$,
2239: then, up to a subsequence,
2240: $$
2241: \si_{\infty,j}= \lim_i \si_{ij}\in W_{aff}
2242: $$
2243: fixes the point $x$. Therefore it follows from the definition of a chain that
2244: $$
2245: \bar\la_{\infty,j} \ge_{W_x} \bar\la_{\infty,j+1}
2246: $$
2247: for each $j$. By putting these inequalities together we obtain
2248: $$
2249: \bar\la_{\infty,1} \ge_{W_x} \bar\la_{\infty,n}. \qed
2250: $$
2251: 
2252: 
2253: 
2254: \medskip
2255: Suppose now that $p_i, q_i$ are Hecke paths,
2256: $$
2257: p_i= \pi_{\la_{i1}} * \pi_{\la_{i2}}*...*\pi_{\la_{in}}, q_i= \pi_{\mu_{i1}} * \pi_{\mu_{i2}}*...*\pi_{\mu_{im}}
2258: $$
2259: and the concatenation $r_i:= p_i*q_i$ satisfies the chain condition. We suppose that $r_{\infty}=p_\infty * q_\infty$
2260: is the limit of the sequence of paths $r_i$ and the paths $p_\infty, q_\infty$ are not constant.
2261: 
2262: 
2263: \begin{lem}
2264: \label{2}
2265: Under the above assumptions, the path $r_\infty$ also satisfies the chain condition.
2266: \end{lem}
2267: \proof Lemma \ref{1} implies that the paths $p_\infty, q_\infty$ are Hecke. Therefore it suffices to verify the chain condition
2268: at the concatenation point $x=p_\infty(1)$. Up to passing to a subsequence, we have:
2269: $$
2270: \lim_{i\to\infty} \la_{ij}=\la_{\infty,j}, \lim_{i\to\infty} \mu_{ij}=\mu_{\infty,j},
2271: $$
2272: $$
2273: \lim_{i\to\infty} \bar\la_{ij}=\bar\la_{\infty,j}, \lim_{i\to\infty} \bar\mu_{ij}=\bar\mu_{\infty,j}.
2274: $$
2275: 
2276: Suppose that
2277: $$
2278: \lim_i \la_{ij}=0, j=k+1,...,n, \quad \lim_i \mu_{ij}=0, j=1,...,l,
2279: $$
2280: and
2281: $$
2282: \la_{\infty,k}\ne 0, \quad \mu_{\infty,l+1}\ne 0.
2283: $$
2284: 
2285: 
2286: By Lemma \ref{1},
2287: $$
2288: \bar\la_{\infty,k} \ge_{W_x} \bar\la_{\infty,n}, \quad \bar\mu_{\infty,1} \ge_{W_x} \bar\mu_{\infty,l+1}.
2289: $$
2290: On the other hand, it is clear that
2291: $$
2292: \bar\la_{\infty,n} \gtrsim_{W_x} \bar\mu_{\infty,1}.
2293: $$
2294: Therefore, by Lemma \ref{ifff}, we get
2295: $$
2296: \bar\la_{\infty,k} \gtrsim_{W_x} \bar\mu_{\infty,l+1}. \qed
2297: $$
2298: 
2299: 
2300: We now can finish the proof Theorem \ref{compactness}. Suppose that $p_i=p_{i1}*...*p_{im}$ is a sequence of paths
2301: in ${\mathcal P}_{chain,m,\eps}$, where each $p_{ij}$ is a billiard path, and $p_{\infty}$ is the limit of this sequence.
2302: Each sequence of billiard paths $(p_{ij})_{i\in \N}$ converges to a billiard path $p_{\infty,j}$ which is a Hecke path according to Lemma \ref{1}.
2303: Consider now a concatenation point $x$ of the subpaths $p_{\infty}:=p_{\infty,j}, q_{\infty}:= p_{\infty,j+1}$. The paths
2304: $p_\infty, q_\infty$ are non-constant by the inequality (\ref{bdd}). Therefore we
2305: can apply Lemma \ref{2} to conclude that the path $p_\infty$ satisfies chain condition at the point $x$. \qed
2306: 
2307: 
2308: 
2309: 
2310: 
2311: 
2312: \begin{rem}
2313: It is easy to see that the assumption that the length of each billiard subpath $p_{ij}$ is bounded away
2314: from zero, is necessary in Theorem \ref{compactness}.
2315: \end{rem}
2316: 
2317: 
2318: 
2319: 
2320: 
2321: 
2322: \section{Folding}
2323: \label{sectionfolding}
2324: 
2325: 
2326: The key tool for proving the main results of this paper is {\em
2327: folding} of polygons in a building $X$ into apartments and Weyl
2328: chambers. The folding construction replaces a geodesic segment $\t{p}$ in $X$
2329: with a piecewise-linear path $p$ in an apartment.
2330: This construction was used in \cite{KLM3} to construct various
2331: counter-examples.  The reader will note that the folding
2332: construction used in the present paper is somewhat different from
2333: the one in \cite{KLM3}.
2334: 
2335: 
2336: 
2337: 
2338: 
2339: 
2340: 
2341: 
2342: \subsection{Folding via retraction}
2343: 
2344: 
2345: Suppose that $X$ is a Euclidean or spherical building modeled on the
2346: Coxeter complex $(A,W)$, we identify the model apartment $A$ with an
2347: apartment $A\subset X$; let $a\subset A$ be an alcove (or a chamber
2348: in the spherical case). Recall that the {\em retraction, or folding,
2349: to an apartment} $f=Fold_{a,A}: X\to A$ is defined as follows (see
2350: for instance \cite{Rousseau}):
2351: 
2352: 
2353: Given a point $x\in X$ choose an apartment $A_x$ containing $x$ and
2354: $a$. Then there exists a (unique) isomorphism $\phi: A_x\to A$
2355: fixing $A\cap A_x$ pointwise and therefore fixing $a$ as well. We
2356: let $f(x):= \phi(x)$. It is easy to see that $f(x)$ does not depend
2357: on the choice of $A_x$. Observe that $f$ is an isometry on each
2358: geodesic $\ol{xy}$, where $y\in a$.
2359: 
2360: 
2361: \medskip
2362: The retraction can be generalized as follows.
2363: 
2364: 
2365: Suppose that $X$ is a Euclidean building, $a$ is an alcove with a
2366: vertex $v$ (not necessarily special). Let $\De\subset A$ denote a
2367: Weyl chamber with  tip $o$. Choose a dilation $h\in Dil(A,W)$ which
2368: sends $v$ to $o$. Let ${\mathbb P}: A\to \De$ denote the natural
2369: projection which sends points $x\in A$ to $W_{sph}\cdot x \cap
2370: \Del$, where $W_{sph}$ is the stabilizer of $o$ in $W_{aff}$.
2371: 
2372: 
2373: We define a folding $g=Fold_{v,h,\De}: X\to \De$ as the composition
2374: $$
2375: \P\circ h\circ Fold_{a,A}.
2376: $$
2377: The mapping $g$ will be called a {\em folding} into a Weyl chamber.
2378: Observe that $g$ (unlike $Fold_{a,A}$) does not depend upon the
2379: choice of the alcove $a$, therefore it will be denoted in what
2380: follows $g=Fold_{v,h,\De}$. Note that
2381: $$
2382: Fold_{v,k\circ h,\De}= k\circ Fold_{v,h,\De}.
2383: $$
2384: 
2385: 
2386: In case when $h=Id$ we will abbreviate $Fold_{v,h,\De}$ to
2387: $Fold_{\De}$.
2388: 
2389: 
2390: 
2391: 
2392: \begin{rem}
2393: The folding maps are Lipschitz and differentiable. The restriction
2394: of the retraction $Fold_{a,A}$ to each chamber (alcove) is a
2395: congruence of two chambers (alcoves).
2396: \end{rem}
2397: 
2398: 
2399: 
2400: 
2401: 
2402: 
2403: %\begin{rem}
2404: %Observe that instead of retracting to an apartment one can retract
2405: %$X$ to the parallel set of a flat in $X$. We will explore this in \cite{HKM}.
2406: %\end{rem}
2407: 
2408: 
2409: If $h=Id$ (and thus $v=o$) one can describe $f=Fold_{\De}$ as
2410: follows. Given a point $x\in X$ find an apartment $A_x$ through $v,
2411: x$ and a Weyl chamber $\De_x\subset A_x$ with  tip $v$. Let
2412: $\phi: \De_x\to \De$ be the unique isometry extending to an
2413: isomorphism of Coxeter complexes $A_x\to A$. Then $f(x)=\phi(x)$.
2414: 
2415: 
2416: \medskip
2417: Suppose now that $X$ is a Euclidean building, $x\in X$; we give the
2418: link $\Si_x(X)$ structure of an unrestricted spherical building $Y$.
2419: Let $R$ denote the corresponding root system. Let $\del$ be a
2420: chamber in $Y$ and $\xi, \mu\in \del$. Let $f: X\to \Del$ be a
2421: folding of $X$ to a Weyl chamber. Let $x'$, $\xi', \mu'$ denote the
2422: images of $x$, $\xi, \mu$ under $f$ and $df_x$. Then
2423: 
2424: 
2425: \begin{lem}
2426: \label{positivity} 1. $d_{ref}(\xi, \mu)= d_{ref}(\xi', \mu')$.
2427: 
2428: 
2429: 2. For each $\al\in R$,
2430: $$
2431: \al(\mu')\ge 0 \iff \al(\xi')\ge 0.
2432: $$
2433: \end{lem}
2434: \proof The restriction $df|_\del$ is an isometry which is the
2435: restriction of an isomorphism of spherical apartments. This proves
2436: (1). To prove (2) observe that $df$ sends $\del$ to a spherical Weyl
2437: chamber $\del'$ in $\Si_{x'}X$. \qed
2438: 
2439: 
2440: \medskip
2441: Let $f$ be a folding of $X$ into an apartment or a chamber.
2442: 
2443: 
2444: \begin{lem}
2445: \label{broken} For each geodesic segment $\ol{xy}\subset X$ its
2446: image $f(\ol{xy})$ is a broken geodesic, i.e. it is a concatenation
2447: of geodesic segments.
2448: \end{lem}
2449: \proof We give a proof in the case of a folding into an apartment
2450: and will leave the other case to the reader. Let $A'\subset X$
2451: denote an apartment containing the geodesic segment $\ol{xy}$. Let
2452: $a_1,...,a_m$ denote the alcoves (or chambers in the spherical case)
2453: in $A'$ covering $\ol{xy}$, set $\ol{x_i x_{i+1}}:= \ol{xy}\cap
2454: a_i$. For each $a_i$ there exists an apartment $A_i$ containing the
2455: alcoves $a$ and $a_i$. The restriction of the retraction $f$ to
2456: $A_i$ is an isometry. It is now clear that the path
2457: $f(\ol{xy})$ is a composition of the geodesic paths $f(\ol{x_i
2458: x_{i+1}})$.
2459:  \qed
2460: 
2461: 
2462: We let $x_i'=f(x_i)$ denote the break points of $f(\ol{xy})$. For
2463: each $x_i'$ let $\xi_i', \eta_i'$ denote the unit tangent vectors in
2464: $T_{x_i'} A$ which are tangent to the segments $\ol{x_i' x_{i-1}'},
2465: \ol{x_i' x_{i+1}'}$ respectively.
2466: 
2467: 
2468: \begin{lem}
2469: \label{billiard}
2470:  The broken geodesic $f(\ol{xy})$ is a {\em billiard
2471: path}, i.e. for each break point $x_i'$ the vectors  $\xi':=\xi_i',
2472: \eta':=\eta_i'$ satisfy
2473: $$
2474: \exists w\in W_{x_i'} : w(\xi')=-\eta'.
2475: $$
2476: \end{lem}
2477: \proof We again present a proof only in the case of a folding into
2478: an apartment. Let $Y=\Si_{x_i} X$ denote the spherical building
2479: which is the space of directions of $X$  at $x_i$. Let $\De_Y$
2480: denote the Weyl chamber of this building and $\theta: Y\to \De_Y$
2481: the canonical projection. The directions $\xi=\xi_i$ and
2482: $\eta=\eta_i$ of the segments $\ol{x_i x_{i-1}}, \ol{x_i x_{i+1}}$
2483: are antipodal in the building $Y$. Since the folding $f$ is an
2484: isomorphism of the apartments $A_i\to A, A_{i+1}\to A$, and
2485: $$
2486: df(\eta)=\eta', \quad df(\xi)=\xi',
2487: $$
2488: we see that
2489: $$
2490: \theta(\xi)=\theta(\xi'), \quad \theta(\eta)=\theta(\eta').
2491: $$
2492: The assertion now follows from Lemma  \ref{thetaproperties}, part
2493: (2). \qed
2494: 
2495: 
2496: \begin{lem}
2497: \label{invarianceoflength} Suppose that $X$ is a Euclidean building,
2498: $f=Fold_{a,A}: X\to A$ and $g=Fold_{z,h,\De}$ are foldings to an
2499: apartment and a chamber respectively. Then for each piecewise-linear path $p$ in
2500: $X$ we have:
2501: 
2502: 
2503: 1. $\length(f(p))=\length(p)$.
2504: 
2505: 
2506: 2. $\length(g(p))=k\cdot \length(p)$, where $k>0$ is the conformal
2507: factor of the dilation $h$.
2508: \end{lem}
2509: \proof We will prove the first assertion since the second assertion
2510: is similar. It suffices to give a proof in the case when $p$ is a
2511: billiard path. Then, analogously to the proof of Lemma
2512: \ref{billiard}, there exists a representation of $p$ as a composition
2513: of geodesic subpaths
2514: $$
2515: p=p_1\cup ...\cup p_m
2516: $$
2517: so that the restriction of $f$ to each $p_i$ is a congruence.
2518: Therefore
2519: $$
2520: \Length(p_i)= \Length(f(p_i))
2521: $$
2522: and hence
2523: $$
2524: \Length (p)= \sum_i \Length(p_i)= \Length (f(p)). \qed
2525: $$
2526: 
2527: 
2528: 
2529: 
2530: 
2531: 
2532: \medskip
2533: {\bf Derivative of the retraction.} We assume that $rank(X)\ge 1$.
2534: We identify the model apartment $A$ with an apartment in $X$. Pick
2535: $a\subset A$ which is an alcove (in the Euclidean case) or a chamber
2536: (in the spherical case). Given a point $x'\in X$ choose an apartment
2537: $(A',W')$ through $a$ and $x'$ and let $\phi: A\to A'$ denote the
2538: inverse to the retraction $f=Fold_{a,A}: A'\to A$. Set $x=f(x')$ and
2539: let $W_x'$ denote the stabilizer of $x'$ in $W'$. Then the link
2540: $Y=\Si_{x'}(X)$  has a natural structure of a thick spherical
2541: building modeled on $(S,W_x')$. It is easy to see that $(S, W_x')$
2542: is independent of the choice of $A'$. Observe that if $x'$ is
2543: antipodal to a regular point $y\in a$ then $x'$ is regular itself
2544: and therefore $W_x'=\{1\}$. We next define a chamber $s\subset S$:
2545: 
2546: 
2547: Given a regular point $y\in a\setminus \{x'\}$ and a geodesic
2548: segment $\ol{x'y}$, let $\zeta=\zeta(y)$ denote the unit tangent
2549: vector to $\ol{x'y}$ at $x'$. Then the set
2550: $$
2551: \{\zeta(y): y \hbox{~is a regular point in~} a\},
2552: $$
2553: is contained in a unique spherical chamber $s\subset S$. (If $x'$ is
2554: antipodal to some $y\in int(a)$ then $s=S$.)
2555: 
2556: 
2557: Set $f':= \phi\circ f= Fold_{a,A'}$.
2558: 
2559: 
2560: 
2561: 
2562: \begin{lemma}
2563: \label{consistent}
2564:  The derivative $d_{x'}f': Y\to S$ equals $Fold_{s,S}$.
2565: \end{lemma}
2566: \proof Given $\eta\in Y$, find an alcove (or a spherical chamber)
2567: $c$ so that $\eta\in \Si_{x'} c$. Then there exists an apartment
2568: $A_\eta\subset X$ containing both $a$ and $c$. Let $S_\eta$ denote
2569: the unit tangent sphere of $A_\eta$ at $x'$. Then $\eta\in S_\eta$
2570: and $s\subset S_\eta$. Now it is clear from the definition that
2571: $$
2572: df'(\eta)=Fold_{s,S}(\eta)$$
2573:  since both maps send $S_\eta$ to $S$ and fix $s$ pointwise. \qed
2574: 
2575: 
2576: 
2577: 
2578: \medskip
2579: {\bf Folding of polygons.} Suppose now that $X$ is a building and
2580: $\t{P}=[\t{z}, \t{x}_1, \ldots, \t{x}_n]$ is a geodesic polygon in $X$. Pick an
2581: apartment $A\subset X$ which contains $\ol{\t{z} \t{x}_1}$ and an alcove
2582: $a\subset A$ which contains $\t{z}$. Let $\De\subset A$ denote a Weyl
2583: chamber (in case $X$ is Euclidean) with  tip $o$. Let $f$ be a
2584: folding of $X$ of the form
2585: $$
2586: f=Fold_{a,A}$$
2587:  or
2588: $$
2589: f=Fold_{\t{z},h,\De},$$
2590:  where $h$ is a dilation sending $\t{z}$ to $o$. We will
2591: then apply $f$ to $\t{P}$ to obtain a {\em folded polygon} $P:= f(\t{P})$
2592: in $A$ or $\De$ respectively.
2593: 
2594: 
2595: Observe that the restriction of $f$ to the edges $\ol{\t{z} \t{x}_1}$ and
2596: $\ol{\t{x}_n \t{z}}$ of $P$ is an isometry or a similarity. The restriction
2597: of $f$ to the path
2598: $$
2599: \t{p}=\ol{\t{x}_1 \t{x}_2}\cup ...\cup \ol{\t{x}_{n-1}\t{x}_n}
2600: $$
2601: preserves the {\em type} of the unit tangent vectors, cf. Lemma
2602: \ref{invarianceoflength}. We will be using foldings into apartments
2603: and chambers to transform geodesic polygons in $X$ into {\em folded
2604: polygons}.
2605: 
2606: 
2607: In the special case when $\t{P}=T$ is a triangle (and thus $n=2$), the
2608: folded triangle $P=f(T)$ has two geodesic sides $\ol{z w_1}:=f(\ol{\t{z} \t{x}_1}),
2609: \ol{x_2 z}:= f(\ol{\t{x}_2 \t{z} })$ and one {\em broken side} $p:=f(\ol{\t{x}_1 \t{x}_2})$, so we
2610: will think of $f(T)$ as a {\em broken triangle}.
2611: 
2612: 
2613: 
2614: 
2615: The next proposition relates folding into a Weyl chamber with the
2616: concept of folding of polygons used in \cite{KLM3}. Let $P=[o,
2617: x_1, x_2, \ldots , x_n]$ be a polygon in $\De$. Triangulate $P$
2618: from the vertex $o$ into geodesic triangles $T_i=[x_i,
2619: x_{i+1},o]$. Suppose that $\t{P}\subset X$ is a geodesic polygon
2620: $$
2621: \t{P}=[o, \t{x}_1, \t{x}_2, \ldots, \t{x}_n], \t{x}_1=x_1,
2622: $$
2623: triangulated into geodesic triangles $\t{T}_i=[\t{x}_i, \t{x}_{i+1},o]$, where
2624: each $\t{T}_i$ is contained in an apartment $A_i$. Assume that for each
2625: $i$ there exists a congruence
2626: $$
2627: \phi_i: \t{T}_i\to T_i
2628: $$
2629: i.e. an isometry sending $\t{x}_j$ to $x_j$ ($j=i, i+1$) which extends
2630: to an isomorphism of Coxeter complexes $\phi_i: A_i\to A$.
2631: 
2632: 
2633: \begin{prop}
2634: \label{uniqueness}
2635:  Under the above assumptions, for each $i$,
2636: $Fold_{\De}|\t{T}_i=\phi_i|\t{T}_i$.
2637: \end{prop}
2638: \proof Let $\De_i\subset A_i$ denote the preimage of $\De$ under
2639: $\phi_i$. Then each $\De_i$ is a Weyl chamber, hence
2640: $\phi_i|\De_i=Fold_{\De}|\De_i$, by the alternative description of
2641: $Fold_{\De}$ given earlier in this section. \qed
2642: 
2643: 
2644: \medskip
2645: The following lemma shows that unfolding of polygons is a {\em local
2646: problem}. Suppose that $T=[o, x_1,...,x_n]\subset A$ is a
2647: geodesic polygon so that $x_i\ne o$ for each $i$. For each $i=2,...,
2648: n-1$ we define the unit vectors
2649: $$
2650: \xi_i, \eta_i, \zeta_i\in \Si_{x_i} A
2651: $$
2652: which are tangent to the segments $\ol{x_i x_{i-1}},
2653: \ol{x_{i}x_{i+1}}$, $\ol{x_i o}$. Define thick spherical
2654: buildings $Y_i:= \Si_{x_i}(X)$. By combining the above proposition
2655: with \cite[Condition 7.5]{KLM3} we obtain
2656: 
2657: 
2658: \begin{lem}
2659: \label{locality} The polygon $T$ can be unfolded in $X$ to a
2660: geodesic triangle $\t{T}$ whose vertices project to $o, x_1, x_n$ if
2661: and only if for each $i=2,...,n-1$, there exists a triangle
2662: $[\t\xi_i, \t\zeta_i, \t\eta_i]\subset Y_i$ so that
2663: $$
2664: d_{ref}(\t\xi_i, \t\zeta_i) = d_{ref}(\xi_i, \zeta_i),
2665: $$
2666: $$
2667: d_{ref}(\t\eta_i, \t\zeta_i) = d_{ref}(\eta_i, \zeta_i),
2668: $$
2669: $$
2670: d(\t\xi_i, \t\eta_i) = \pi.
2671: $$
2672: \end{lem}
2673: 
2674: 
2675: 
2676: 
2677: \medskip
2678: We will eventually obtain a characterization of the broken triangles
2679: in $\De$ which are foldings of geodesic triangles in $X$ as {\em
2680: billiard triangles satisfying the chain condition}, see section
2681: \ref{last}. The goal of the next section is to give a {\em
2682: necessary} condition for a broken triangle to be unfolded; we also
2683: give a partial converse to this result.
2684: 
2685: 
2686: 
2687: 
2688: 
2689: 
2690: \subsection{Converting folded triangles in spherical buildings into chains}
2691: \label{converting}
2692: 
2693: Suppose that $X$ is a (thick) spherical or Euclidean building modeled on $(A,W)$. Consider a
2694: triangle $\tilde{T}=[x, y, z]\subset X$ with
2695: $$
2696: \be:=d_{ref}(x, y), \quad \ga:=d_{ref}(y, z).
2697: $$
2698:  Assume that $A$ is embedded in $X$ so that it contains $x$ and $z$.
2699: Let $a\subset A$ be a spherical chamber or a Euclidean alcove containing $z$.
2700: In the spherical case we regard $a$ as the {\em negative chamber} in $A$,
2701: let $\De$ denote the positive chamber $-a$. We have the retraction $f:=
2702: Fold_{a,A}:X\to A$.
2703: 
2704: 
2705: 
2706: 
2707: 
2708: 
2709: \begin{figure}[tbh]
2710: \centerline{\epsfxsize=5in \epsfbox{chain.eps}} \caption{\sl
2711: Converting geodesic triangle to a chain.} \label{chain.fig}
2712: \end{figure}
2713: 
2714: 
2715: 
2716: 
2717: \begin{thm}\label{T2}
2718: There exists a $(A,W,a)$--chain $(y_0,..., y_m)$  such that
2719: $y_m=f(y)$, $d_{ref}(x, y_0)=\be$, $d_{ref}(y_m, z)=\ga$. (In the case when $X$ is a Euclidean building
2720: the above chain is an affine chain.)
2721: See Figure \ref{chain.fig}.
2722: \end{thm}
2723: \proof
2724: We prove the assertion for the spherical buildings as the Euclidean case is completely analogous.
2725: (This is also the only case when this theorem is used in the present paper.)
2726: 
2727: Our proof is by induction on the rank of the building.
2728: Consider first the case when $rank(X)=0$ (i.e. $A=S^0$ is the
2729: 2-point set). If $y$ and $x$ are both distinct from $z$, then
2730: $f(y)\ne z$. This implies that $f(y)=x$ and we take
2731: $$
2732: y_0:= z, y_1:= y, m=1.
2733: $$
2734: In the remaining cases we will use the chain $y_0=f(y)=y_m$.
2735: 
2736: 
2737: Suppose now that $rank(X)=r\ge 1$ and the assertion holds for all
2738: (spherical) buildings of rank $r-1$, let's prove it for buildings of
2739: rank $r$.
2740: 
2741: 
2742: We let $\tilde{p}: [0, c]\to \ol{xy}$ denote the unit speed parametrization
2743: of $\ol{xy}$ and set $p:= f(\t{p})$. {\em We assume for now that $z\notin \ol{xy}$}.
2744: 
2745: 
2746: As in the proof of Lemma \ref{broken}, we ``triangulate'' the geodesic
2747: triangle $\t{T}$ into geodesic triangles $\t{T}_i:= [z, \t{x}_i,
2748: \t{x}_{i+1}]$, where the points $\t{x}_i=\t{p}(t_i)$, $i=1,...,n$, are chosen
2749: so that each triangle $\t{T}_i$ is contained in an apartment
2750: $A_i\subset X$ and the map $f$ restricts to an isometry $f:
2751: \t{T}_i\to f(\t{T}_i)\subset A$. Here $\t{x}_0:=x, \t{x}_{n+1}=y$.
2752: Observe that each side of $\t{T}_i$ has positive length.
2753: 
2754: 
2755: 
2756: 
2757: Let $\t{S}_i:=\Si_{\t{x}_i}(A_i)$ denote the unit tangent sphere at
2758: $\t{x}_i$. Define $\t{s}_i$ to be the (unique) chamber in $\t{S}_i$
2759: containing all the directions of the geodesic segments from
2760: $\t{x}_i$ to the interior of $a$. This determines the {\em positive
2761: chamber} $\t{\De}_i=-\t{s}_i\subset S_i$.
2762: 
2763: 
2764: Set
2765: $$
2766: \t\eta_i:=\t{p}'(t_i), \t\xi_i:= -\eta_i,
2767: $$
2768: and let $\t\zeta_i\in \t{S}_i$ denote the unit tangent vector to
2769: $\ol{\t{x}_i z}$.
2770: 
2771: 
2772: Now, applying the retraction $f$ to all this data, we obtain:
2773: 
2774: 1. The folded triangle $T=f(\t{T})$ which has two geodesic sides
2775: $\ol{zx}, \ol{zy_m}$ (where $y_m=f(y)$), and the broken side
2776: represented by the path $p=f(\t{p})$. In particular, $d_{ref}(y_m, z)=d_{ref}(y, z)$ (as required by the theorem).
2777: 
2778: 
2779: 2. The vertices $x_i=p(t_i)= f(\t{x}_i)$ of the broken geodesic $p$.
2780: 
2781: 3. Unit tangent vectors $\xi_i=df(\t{\xi}_i), \eta_i=df(\t{\eta}_i),
2782: \zeta_i=df(\t{\zeta_i})$ in $\Si_x A$. These vectors are tangent to
2783: the segments $\ol{x_i x_{i-1}}, \ol{x_i x_{i+1}}, \ol{x_i z}$
2784: respectively.
2785: 
2786: 
2787: 4. The positive chamber $\De_i=df(\t{\De}_i)$ and the negative
2788: chamber $s_i=df(\t{s}_i)$ in the spherical Coxeter complex
2789: $(S_i=\Si_{x_i}(A), W_i=W_{x_i})$. The negative chamber contains the
2790: directions tangent to the geodesic segments from $x_i$ to the
2791: chamber $a\subset A$.
2792: 
2793: Our goal is to convert the broken side $p$ of $T$ into a chain in
2794: $A$ by ``unbending'' the broken geodesic $p$ to a geodesic segment
2795: in $A$. See Figure \ref{con.fig}.
2796: 
2797: 
2798: 
2799: \begin{figure}[tbh]
2800: \centerline{\epsfxsize=5in \epsfbox{cone.eps}} \caption{\sl
2801: Forming a chain by unbending.} \label{con.fig}
2802: \end{figure}
2803: 
2804: 
2805: \begin{lem}
2806: \label{trivial} The path $p$ satisfies the following:
2807: 
2808: 
2809: 1. The metric lengths of $p$ and $\tilde{p}=\ol{xy}$ are the same.
2810: 
2811: 
2812: 2. $p'(0)=\tilde{p}'(0)$.
2813: 
2814: 
2815: 3. At each break-point $x_i$ there exists an $(S_i, W_{i}, s_i)$-chain from $-\xi_i$ to $\eta_i$.
2816: \end{lem}
2817: \proof The first two assertions are clear from the construction.
2818: Let's prove the last statement. For each $i$ and the point
2819: $\t{v}=\t{x}_i$ we have the spherical building $Y:=\Si_{\t{v}}(X)$
2820: which has rank $r-1$. This building contains the antipodal points
2821: $$
2822: \t\xi_i, \t\eta_i
2823: $$
2824: and the point $\t\zeta_i$. We form the geodesic triangle
2825: $\tau=[\t\xi_i, \t\eta_i, \t\zeta_i]\subset Y$, where we use an
2826: arbitrary shortest geodesic in $Y$ to connect $\xi_i$ to $\eta_i$.
2827: Therefore $\xi_i, \eta_i, \zeta_i$ are vertices of the broken
2828: geodesic triangle $df(\tau)\subset S_i$.
2829: 
2830: 
2831: As in Lemma \ref{consistent}, we use the isomorphism $S_i\to
2832: \t{S}_i$ (sending $s_i$ to $\t{s}_i$) to identify these apartments.
2833: Under this identification, $df: Y\to S_i$ is the retraction
2834: $Fold_{s_i, S_i}$ of $Y$ to the apartment $S_i$. Thus $df(\tau)$ is
2835: a folded triangle in $S_i$.
2836: 
2837: 
2838: Hence, by the (rank) induction hypothesis, for each $i$ there exists
2839: a chain
2840: $$
2841: (-\xi_i,..., \eta_i)
2842: $$
2843: in the spherical Coxeter complex $(S_i, W_{i}, s_i)$. \qed
2844: 
2845: 
2846: 
2847: 
2848: \begin{lem}
2849: For each path $p: [0, c]\to A$ satisfying the conclusion of Lemma
2850: \ref{trivial}, there exists a point $y'\in A$ such that
2851: $$
2852: d_{ref}(x, y')= \be %= \Length(p),
2853: $$
2854: and
2855: $$
2856: y' \ge p(c)
2857: $$
2858: in $(A,W,-\De)$.
2859: \end{lem}
2860: \proof We use the second induction, on the number $n$ of vertices in
2861: the broken geodesic $p$. Set $u:= p(c)$.
2862: 
2863: 
2864: The metric length of the path $p$ equals the metric length of the
2865: path $\t{p}=\ol{x y}$, the tangent directions of these paths at $x$
2866: are the same. Therefore, if $n=0$ (and hence the path $p$ is
2867: geodesic) there is nothing to prove, one can simply take $y'=u$.
2868: 
2869: 
2870: Assume that the assertion holds for all $n\le N-1$, let's prove it
2871: for $N$. We treat the path $p$ as the composition
2872: $$
2873: p|[0, t_N]\cup \ol{x_N u}
2874: $$
2875: Our goal is to replace the geodesic subpath $\ol{x_N u}$ with a
2876: geodesic path $w( \ol{x_N u})$, where $w\in W$ is fixing $x_N$, so
2877: that:
2878: 
2879: 
2880: 1. $\ol{x_{N-1} x_N}\cup w( \ol{x_N u})$ is a geodesic segment.
2881: 
2882: 
2883: 2. There exists an $(A, W, -\De)$-chain between $w(u)$ and $u$. Then
2884: we would be done by the induction on $n$. Indeed, the new path
2885: $$
2886: p|[0, t_N]\cup w(\ol{x_N u})
2887: $$
2888: has one less break-point and still satisfies the conclusion of Lemma
2889: \ref{trivial}. Thus, by the induction hypothesis, there exists
2890: $y'\in A$ so that
2891: $$
2892: y' \ge w(u) \ge u \Rightarrow y' \ge u,
2893: $$
2894: $$
2895: d_{ref}(x, y')= d_{ref}(x, y).
2896: $$
2897: 
2898: 
2899: {\bf Construction of $w$.} Recall that there exists an $(S_N, W_{N},
2900: -\De_N)$-chain
2901: $$
2902: (-\xi_N=\nu_0,..., \nu_k=\eta_N),
2903: $$
2904: hence we have a sequence of reflections  $r_1,...,r_k\in W_{N}$
2905: (fixing walls $H_i\subset S_N$, $i=1,...,k$) so that:
2906: $$
2907: r_i(\nu_{i-1})=\nu_i, i=1,...,k,
2908: $$
2909: and each wall $H_i$ separates $\nu_{i}$ from the negative chamber
2910: $s_N$. We extend each reflection $r_i$ from $S_N$ to a reflection
2911: $r_i$ in $A$, and each $H_i$ to a wall $H_i$ in $A$.
2912: 
2913: 
2914: We therefore define the following points in $A$:
2915: $$
2916: y_k:=u, y_{k-1}:= r_{k}(y_k), y_{k-2}:= r_{k-1}(y_{k-1}),\ldots,
2917: y_{0}:= r_{1}(y_1).
2918: $$
2919: Note that the directions $\nu_i$ are tangent to the segments
2920: $\ol{x_N y_i}$. Thus for each $i$, the wall $H_i$ separates the
2921: point $y_{i}$ from the negative chamber $a\subset A$ and the
2922: sequence
2923: $$
2924: (y_0,...,y_k=u)
2925: $$
2926: forms a chain. We set $w=r_1\circ ... \circ r_k$. The vector $\nu_0$
2927: is antipodal to $-\xi_N$, hence the path
2928: $$
2929: p|[0, T_N] \cup w(\ol{x_N u})
2930: $$
2931: is geodesic at the point $x_N=p(T_N)$. \qed
2932: 
2933: 
2934: This concludes the proof of Theorem in the case when $z\notin\ol{x
2935: y}$.
2936: 
2937: 
2938: We now consider the special case when the above proof has to be
2939: modified: The triangle $T$ is degenerate, i.e. $z\in \ol{xy}$, but
2940: the alcove $a$ is such that $y\notin A$.  Thus the folding
2941: $Fold_{a,A}$ is not an isometry on $T$. Then the tangent direction
2942: $\t\zeta_i$ is not defined when $x_i=z$. Note that $x_i=z$ then is
2943: the only break-point in the broken side of $T'$.
2944: 
2945: 
2946: In this case we replace the vertex $z$ with an arbitrary point $z'$
2947: in the interior of $a$ and repeat the above arguments. The chains
2948: constructed in the process will be independent of the choice of $z'$
2949: and thus, after taking the limit $z'\to z$, we obtain a chain as
2950: required by the assertion of Theorem. \qed
2951: 
2952: 
2953: 
2954: 
2955: \begin{cor}
2956: Cf. \cite[Theorem 8.2, Part 4]{KLM3}.  Suppose that $X$ is a Euclidean building.
2957: Assume that $\al:=d_\De(z,x), \be:=d_\De(x,y), \ga:=d_\De(y,z)$ are in $P(R^\vee)$
2958: and $x, y, z$ are special vertices of $X$. Then
2959: $$
2960: \al+\be+\ga \in Q(R^\vee).
2961: $$
2962: \end{cor}
2963: \proof
2964: Let $(y_0,...,y_m)$ be an affine chain given by Theorem \ref{T2}.
2965: We regard the point $x$ as the origin $o$ in $A$; thus we will regard $z$, $y_0, y_m$ as vectors in $V$.
2966: Then, according to Lemma \ref{difference},
2967: $$
2968: y_m-y_0\in Q(R^\vee).
2969: $$
2970: Consider the vectors $\be':= y_0-x, \ga':= z-y_m, \al':= x-z$ in $P(R^\vee)$. By the definition of $\De$-length,
2971: $$
2972: \al'\in W_{sph} \al, \quad \be'\in W_{sph} \be, \quad \ga'\in W_{sph} \ga.$$
2973:  Therefore, by applying  Lemma \ref{difference} again we see that
2974: the differences
2975: $$
2976: \al-\al',  \quad \be-\be',  \quad \ga-\ga'
2977: $$
2978: all belong to $Q(R^\vee)$. Since
2979: $$
2980: \al'+\be'+\ga'=y_0-y_m\in Q(R^\vee),
2981: $$
2982: the assertion of lemma follows. \qed
2983: 
2984: 
2985: 
2986: 
2987: 
2988: \medskip
2989: The following simple proposition establishes a partial converse to
2990: Theorem \ref{T2}:
2991: 
2992: 
2993: \begin{prop}
2994: \label{simplechain} Suppose that $X$ is a thick spherical building and, as before, the point $z$ belongs to a
2995: negative chamber $a=-\De$. Then, for each simple chain $(y_0, y_1)$
2996: such that $d_{ref}(x, y_0)=\pi, d_{ref}(z, y_1)=\be$, there exists a
2997: point $y\in X$ so that
2998: $$
2999: d_{ref}(y, z)= \ga \quad \hbox{and}\quad  d_{ref}(x , y)= \pi.$$
3000: \end{prop}
3001: 
3002: \begin{rem}
3003: Recall that $d_{ref}(x , y)= \pi$ means that the points $x$ and $y$ are antipodal.
3004: \end{rem}
3005: 
3006: \proof Let $\tau(y_0)=y_1$ where $\tau$ is a reflection in a wall
3007: $H\subset A$ as in the definition of a chain. Let $A=A^-\cup A^+$ be
3008: the union of half-apartments, where $A^-$ is bounded by $H$ and contains $a$.
3009: By the definition of a chain, $y_1\in A^-, y_0\in A^+$
3010: and hence the antipodal point $x=-y_0$ belongs to $A^+$.
3011: 
3012: 
3013: Since $X$ is thick, there exists a half-apartment $B^-\subset X$
3014: which intersects $A$ along $H$. Define the apartment $B:= A^-\cup
3015: B^-$; then there exists  an isomorphism of Coxeter complexes
3016: $$
3017: \phi: A\to B, \phi|A^-=id.
3018: $$
3019: We set $y:= \phi(y_0)$.
3020: 
3021: 
3022: Since $\phi$ is an isomorphism of Coxeter complexes which fixes $z$,
3023: it preserves the refined distance to the point $z$ and hence
3024: $$
3025: d_{ref}(y, z)= d_{ref}(y_1, z)=\ga.$$
3026: 
3027: 
3028: The union $C:= A^+\cup B^-$ is also an apartment in $X$. Then there
3029: exists an isomorphism  $\psi: B^-\to A^-$ so that
3030: $$
3031: \psi\circ\phi|A^+= \tau|A^+.
3032: $$
3033: The isomorphism $\psi$ extends to an isomorphism $\rho: C\to A$
3034: fixing $A^+$ pointwise and hence fixing the point $x$. Therefore
3035: $$
3036: d_{ref}(x , y)= d_{ref}(x, y_0)= \pi.  \qed
3037: $$
3038: 
3039: 
3040: 
3041: 
3042: 
3043: \subsection{Folding polygons in Euclidean buildings}
3044: \label{euclideanfolding}
3045: 
3046: 
3047: Our next goal is to show that each folding transforms certain piecewise-linear
3048: paths in Euclidean buildings to paths satisfying the chain condition.
3049: 
3050: 
3051: Suppose that $X$ is a Euclidean building with model apartment
3052: $(A,W_{aff})$, $\De\subset A$ is the positive Weyl chamber with the
3053: tip $o$. Consider a piecewise-linear path $\tilde{p}: [0,c]\to \t{A}$, which is
3054: parameterized  with the unit speed, where $\tilde{A}\subset X$ is an
3055: apartment. We assume that for each $t\in [0,c]$
3056: $$
3057: \tilde{p}'_-(t) \sim_{W_{sph}} \tilde{p}'_+(t),
3058: $$
3059: for instance, $\tilde{p}$ could be a geodesic path.
3060: 
3061: 
3062: Thus the path $\tilde{p}$ {\em trivially satisfies the chain
3063: condition}. Let $g: X\to \De$ be a folding into $\De$,
3064: $g=Fold_{z,h,\De}$ for a certain $z\in A$ and $h$. Recall that the
3065: folding $g$ is the composition of three maps:
3066: $$
3067: g= \P_{\Del} \circ h \circ f, \quad f=Fold_{a,A},
3068: $$
3069: where $a$ is an alcove in $A$ containing $z$, $h\in Dil(A,W_{aff})$ is a
3070: dilation sending $z$ to the point $o$. Consider
3071: the structure of a Coxeter complex on $A$ given by the pull-back
3072: $$
3073: h^*(A,W_{aff}).
3074: $$
3075: We thus get a new (typically non-thick) building structure for $X$,
3076: the one modeled on $h^*(A,W_{aff})$.
3077: 
3078: \begin{defn}
3079: We say that a path $\tilde{p}$ is {\em generic} if it is disjoint
3080: from $z$ and from the codimension 2 skeleton of $X$ and the
3081: break-points of $\tilde{p}$ are disjoint from the codimension 1
3082: skeleton of $X$, where $X$ is regarded as a building modeled on
3083: $h^*(A,W_{aff})$.
3084: \end{defn}
3085: 
3086: 
3087: 
3088: The main result of this section is
3089: 
3090: 
3091: \begin{thm}
3092: \label{chaincondition} The folded path $p= g(\tilde{p})$ satisfies
3093: the chain condition.
3094: \end{thm}
3095: \proof The proof of this theorem is mostly similar (except for the
3096: projection $\P$ which causes extra complications) to the proof of
3097: Theorem \ref{T2} in the previous section. We will prove Theorem
3098: \ref{chaincondition} in two steps: We first establish it for the
3099: paths $\tilde{p}$ which are {\em generic}. Then we use the {\em
3100: compactness theorem} to prove it in general.
3101: 
3102: 
3103: \begin{prop}
3104: \label{chaincondition0} The conclusion of Theorem
3105: \ref{chaincondition} holds for {\em generic} paths $\tilde{p}$.
3106: \end{prop}
3107: \proof If a point $\tilde{x}=\tilde{p}(t)$ is a regular point of
3108: $X$, then
3109: $$
3110: dg_{\tilde{x}}: \Si_{\tilde{x}}(X)\to \Si_x(A), \quad x=g(\tilde{x})
3111: $$
3112: is an isometry. Thus the path $p$ trivially satisfies the chain
3113: condition at the point $x$.
3114: 
3115: 
3116: Therefore we assume that $\tilde{x}$ is a singular point of $X$. Since
3117: $\tilde{p}$ is assumed to be generic, this point lies on exactly one
3118: wall of $X$; moreover, $\tilde{p}$ is geodesic near $\tilde{x}$.
3119: 
3120: 
3121: 
3122: 
3123: We first analyze what happens to the germ of $\tilde{p}$ at
3124: $\tilde{x}$ under the retraction $f$. We suppose that the
3125: restriction of $f$ to the germ $(\t{p}, \t{x})$ is not an isometry
3126: (otherwise there is nothing to discuss).  Let $\tilde\zeta\in
3127: \Si_{\tilde{x}}(X)$ denote the tangent to the geodesic
3128: segment $\ol{\tilde{x}z}$. Let $\eta\in \Si_{\tilde{x}}(X\cap
3129: \tilde{A})$ be the tangent vector $p'(t)$, $\tilde\xi:= -\tilde\eta$
3130: (this vector is also tangent to the path $p$). Set $\be:=
3131: d_{ref}(\tilde\zeta, \tilde\eta)$.
3132: 
3133: 
3134: We obtain the  triangle $\tau=[\tilde\xi, \tilde\eta,
3135: \tilde\zeta]$ in $\Si_{\tilde{x}}(X)$. The derivative of the
3136: retraction $f$ at $\t{x}$ is a retraction of the spherical building
3137: $\Si_{\t{x}}(X)$ into its apartment $S$, after identification of $S$
3138: with the sphere $S_{x'}(A)$, $x':=f(\t{x})$ (see Lemma
3139: \ref{consistent}). Define the following elements of $S$:
3140: $$
3141: \eta':=df_{\tilde{x}}(\tilde\eta), \quad \xi':=
3142: df_{\tilde{x}}(\tilde\xi), \quad
3143: \zeta':=df_{\tilde{x}}(\tilde\zeta).
3144: $$
3145: 
3146: 
3147: Therefore, according to Theorem \ref{T2}, the folded triangle
3148: $\tau'=df_x(\tau)\subset S$ yields an $(S,W_{x'},
3149: -\De_{x'})$-chain\footnote{Which is necessarily simple since
3150: $\tilde{p}$ is assumed generic.}
3151: $$
3152: (\mu'=-\xi', \eta')
3153: $$
3154: such that $d_{ref}(\zeta, \eta')= d_{ref}(\t\zeta, \t\eta)=\be$.
3155: 
3156: 
3157: Here $-\De_{x'}$ is a chamber in $(S,W_{x'})$ which contains
3158: the unit tangent vector to the segment $\ol{x' z'}$ where $z'\in a$ is a regular point.
3159: 
3160: 
3161: \begin{rem}
3162: Note that our assumptions on $\tilde{p}$ imply that $(S,W_{x'})$
3163: has a unique wall. If the corresponding wall in $A$ does not
3164: pass through $z$, then the negative chamber $-\De_{x'}$  in $(S,W_{x'})$ is
3165: uniquely determined by the condition that it contains the direction
3166: tangent to
3167:  $\ol{ x' z}$.
3168: \end{rem}
3169: 
3170: 
3171: Consider now the effect of the rest of the folding $g$ on the path
3172: $\tilde{p}$ at $\tilde{x}$. Let $x:= g(\tilde{x})$. We identify $S$
3173: with the unit tangent sphere at the point $x$.
3174: 
3175: 
3176: The dilation $h$ clearly preserves the chain condition at $x'$
3177: (since it acts trivially on the unit tangent sphere). The
3178: restriction of the projection $\P=\P_{\Del}$ to the germ of
3179: $hf(\tilde{p})$ at $hf(\tilde{x})$ is necessarily an isometry (since
3180: $\tilde{p}$ is generic), hence it is given by an element $w\in
3181: W=W_{sph}$. This element transforms the above chain to another
3182: $(S,W_x, -\De_x)$--chain, where
3183: $$
3184: \De_x:= d(w\circ h) (\De_{x'}).
3185: $$
3186: What is left to verify is that the positive chamber $\De_x$ in this
3187: complex contains a translate of the positive chamber $\De$. In case
3188: when $x$ belongs to the interior of $\De$, the segment $\ol{ox}$ is
3189: not contained in any wall and thus the negative chamber $-\De_x$ has
3190: to contain the initial direction of the segment $\ol{xz}$ (see the
3191: remark above). However this initial direction belongs to $-\Del$ and
3192: thus $\De_x$ contains $\De$.
3193: 
3194: 
3195: Consider the exceptional case when $x$ is on the boundary of $\De$.
3196: It then belongs to a unique wall $H$ in the Coxeter complex
3197: $(A,W_{aff})$ and this wall passes through the origin $o$. Rather
3198: than trying to use Theorem \ref{T2} to verify the chain condition at
3199: $x$, we give a direct argument. Let $\eta, \xi$ be the unit vectors
3200: which are the images of $\eta', \xi'$ under
3201: $$
3202: d(w\circ h): \Si_{x'}(A)\to \Si_x(A).
3203: $$
3204: Since the path $p$ is entirely contained in $\De$, the vector
3205: $p'_-(t)$ points {\em outside} of $\De$ and the vector $p'_+(t)$
3206: points {\em inside}. The reflection $\si$ in the wall $H$ sends the
3207: vector $-\xi=p'_-(t)$ the vector $\eta=p'_+(t)$. It is then clear
3208: that the (simple) chain condition is satisfied at the point $x$.
3209: 
3210: 
3211: \medskip
3212: Lastly, we consider the points $\tilde{x}=\tilde{p}(t)$ for which
3213: $f$ is an isometry on the germ of $\tilde{p}$ at $\tilde{x}$.
3214: The point $x=g(\tilde{x})$ belongs to
3215: a face of $\De$ contained in a wall $H$, and this is the only wall
3216: of $(A,W_{aff})$ which passes through $x$. Then,
3217: necessarily, the germ of the path $hf(\tilde{p})$ at
3218: $hf(\tilde{x})$ is a geodesic.  We now simply repeat the
3219: arguments of the exceptional case in the above proof (see also the
3220: proof of Proposition \ref{chaincondition1}) to see that $p$
3221: satisfies the chain condition at $x$. \qed
3222: 
3223: 
3224: \medskip
3225: We are now ready to prove Theorem \ref{chaincondition} for arbitrary
3226: paths $\tilde{p}$. We will do so by approximating the path
3227: $\tilde{p}$ via {\em generic paths}. Let $\la$ be an arbitrary
3228: vector in $\t{A}$. We let $\tilde{q}_\la:=\tilde{p}+\la$ denote the
3229: translation of the path $\tilde{p}$ by the vector $\la$. It is
3230: clear, from the dimension count, that for an open and dense set of
3231: vectors $\la$, the path $\tilde{q}_\la$ is {\em generic}.
3232: 
3233: 
3234: Since the folding $g$ is continuous,
3235: $$
3236: p=g(\tilde{p})= \lim_{\la\to 0} g(\tilde{q}_\la).
3237: $$
3238: By the Proposition \ref{chaincondition0}, each $g(\tilde{q}_\la)$
3239: satisfies the chain condition. Observe that the $\De$-lengths of the
3240: paths $\tilde{p}+\la$ are independent of $\la$. Since $f$ and $\P$
3241: preserve $\De$-lengths of piecewise-linear paths and the dilation $h$ changes them
3242: by a fixed amount, we can apply the compactness theorem (Theorem
3243: \ref{compactness}) to conclude that the limiting path $p$ satisfies
3244: the chain condition as well. \qed
3245: 
3246: 
3247: \medskip
3248: We now verify that, at certain points, the folded path $p$ satisfies
3249: the {\em maximal chain condition}.
3250: 
3251: 
3252: 
3253: 
3254: \begin{prop}
3255: \label{chaincondition1} Under the assumptions of Theorem
3256: \ref{chaincondition} let $\tilde{x}=\tilde{p}(t)$ be such that the
3257: folding $f$ restricts to an isometry on the germ $(\tilde{p},
3258: \tilde{x})$. Then the path $p=g(\tilde{p})$ satisfies the {\em
3259: maximal chain condition} at $x=g(\tilde{x})$.
3260: \end{prop}
3261: \proof Our proof follows Littelmann's arguments in his proof of the
3262: PRV Conjecture  \cite{Littelmann1}. We fold the path
3263: $q:=hf(\tilde{p})$ into $\De$ inductively.
3264: 
3265: 
3266: We subdivide the interval $[0,c]$ as
3267: $$
3268: 0=t_0 < t_1<...< t_k=c
3269: $$
3270: such that $[t_i, t_{i+1}]$ are maximal subintervals so that $q|[t_i,
3271: t_{i+1}]$ is contained in a Weyl chamber of $W_{sph}$.
3272: 
3273: 
3274: We first apply to $q$ an element $w_0\in W_{sph}$ which sends $q([0,
3275: t_1])$ into $\De$, so we can assume that this subpath belongs to
3276: $\De$. Assume that the restriction of $f$ to the germ $(\t{p}, \t{p}(t_1))$ is an isometry.
3277: Let $\mu', \eta'$ be the vectors $q'_-(t_1), q'_+(t_1)$. Then
3278: $$
3279: \mu'\sim \eta',
3280: $$
3281: see Lemma \ref{positivity}. Set $x:= q(t_1)$. The image $\eta$ of
3282: the vector $\eta'$ under $\P$ is obtained as
3283: $$
3284: dw_1(\eta'),
3285: $$
3286: where $w_1\in W_{sph}$ fixes the point $x$ and $\eta$ is the unique
3287: vector in the $W_x$-orbit of $\eta'\in S_{x}$ which points inside
3288: $\De$. Below we describe $w_1$ as a composition of reflections.
3289: 
3290: 
3291: Let $R'$ denote the root subsystem in $R$ generated by the set of
3292: simple roots $\Phi'$ which vanish at the point $x$. Let $\De'$
3293: denote the positive chamber for $W_x$ defined via $\Phi'$. Then the
3294: vector $\eta$ can be described as the unique vector in the
3295: $W_x$-orbit of $\eta'$ (now, regarded as a vector in $V=T_o(A)$)
3296: which belongs to the interior of $\De'$. According to Lemma
3297: \ref{simplesubsystem},
3298: $$
3299: w_1= \tau_m\circ .. \circ \tau_1, \hbox{~where for each~} i, \quad
3300: \tau_i=\tau_{\be_i}, \be_i\in \Phi',
3301: $$
3302: so that the sequence of vectors
3303: $$
3304: (\eta_0=\eta', \eta_1:=
3305: \tau_1(\eta_{0}),...,\eta_m=\tau_m(\eta_{m-1})=\eta),$$ is a chain
3306: in $(S, W_x,-\De)$ which is maximal as a chain in $(S,W_{sph},-\De)$.
3307: 
3308: 
3309: We therefore apply the identity transformation to the path $q|[0,
3310: t_1]$ and the element $w_1$ to the path $q|[t_1, c]$ to transform
3311: the path $q$ to the new path
3312: $$
3313: q_1= q|[0, t_1] \cup w_1\circ q|[t_1, c].
3314: $$
3315: Clearly, $\P(q)= \P(q_1)$ and $\eta$ is the unit vector tangent to
3316: $q_1|[t_1,c]$ at $x$. The above arguments therefore show that $q_1$
3317: satisfies the {\em maximal chain condition} at the point $x$.
3318: 
3319: 
3320: We then proceed to the next point $t_2$, $q_1(t_2)$ belongs to the
3321: boundary of $\De$ and we transform $q_1$ to $q_2$ by
3322: $$
3323: q_2|[0,t_2]=q_1|[0,t_2], \quad q_2|[t_2, c]= w_2\circ q_1|[t_2,c],
3324: $$
3325: where $w_2$ is a certain element of $W_{sph}$ fixing
3326: $q_1(t_2)=q_2(t_2)$. Therefore $\P(q_2)=\P(q_1)=\P(q)$ and we repeat the
3327: above argument. \qed
3328: 
3329: 
3330: \begin{defn}
3331: Suppose that $P=\ol{zx}\cup p \cup \ol{yz}$ is a polygon in $A$,
3332: where $p:[0,1]\to A$ is a piecewise-linear path such that $p(0)=x, p(1)=y$. We say
3333: that $P$ {\em satisfies the chain condition} (resp. simple chain
3334: condition, resp. maximal chain condition) if its subpath $p$
3335: satisfies the chain condition (resp. simple chain condition, resp.
3336: maximal chain condition).
3337: \end{defn}
3338: 
3339: 
3340: Therefore, as an application of Theorem \ref{chaincondition} we
3341: obtain
3342: 
3343: 
3344: 
3345: 
3346: \begin{cor}
3347: \label{folded-chain}
3348:  Suppose that $T=[\t{z}, \t{x}, \t{y}]\subset X$ is a
3349: geodesic triangle, $\t{z}$ is a special vertex which belongs to an
3350: alcove $a\subset A$. Let $\De\subset A$ be a Weyl chamber with the
3351: tip $\t{z}=o$ and $P=Fold_{\De}(T)$ be the folding of $T$ into $\De$. Then
3352: the folded triangle $P$ satisfies the chain condition.
3353: \end{cor}
3354: 
3355: 
3356: 
3357: 
3358: A converse to this corollary will be proven in Theorem
3359: \ref{characterization}; the following is a {\em partial} converse to
3360: Corollary \ref{folded-chain} (which is essentially contained in
3361: \cite[Lemma 7.7]{KLM3}):
3362: 
3363: 
3364: \begin{cor}
3365: \label{chain-folded0} Let $\De\subset A$ be a Weyl chamber with  tip
3366: $o$ in $X$. Suppose that a polygon $P= [o, x_1,...,x_{n}]\subset
3367: \De$, satisfies the {\em simple chain} condition (at each vertex
3368: $x_i, 0<i<n$) and
3369: $$
3370: p=\ol{x_1 x_2}\cup ... \cup \ol{x_{n-1} x_n}
3371: $$
3372: is a {\em billiard} path. Then $P$ unfolds to a geodesic triangle
3373: $T\subset X$, i.e. $Fold_{\De}(T)=P$.
3374: \end{cor}
3375: \proof Let $f:=Fold_{\De}$.  We run the argument from the proof of
3376: Theorem \ref{T2} in the reverse; the reader will observe that our
3377: argument is essentially the same as in the proof of the Transfer
3378: Theorem in \cite{KLM2}. Triangulating the polygon $P$ from the
3379: vertex $o$ we obtain geodesic triangles $P_i=[o, x_i, x_{i+1}]$, $i=1, 2,...,n-1$.
3380: Let $\xi_i, \zeta_i, \eta_i \in \Si_{x_i}(A)$ denote the unit tangent vectors to the
3381: segments $\ol{x_i x_{i-1}}, \ol{x_i o}, \ol{x_i x_{i+1}}$
3382: respectively.
3383: 
3384: We unfold $P$ inductively. Set $T_1:= P_1$; let $A_1:= A$,
3385: this apartment contains the triangle $T_1$. Set $\t{x}_1:= x_1, \t{x}_2:= x_2$,
3386: 
3387: Suppose that we have constructed apartments $A_i\subset X$ and flat triangles
3388: $T_i=[o, \t{x}_i, \t{x}_{i+1}]\subset A_i$, $i=1,...,m-1$,
3389: so that $T_i$ is congruent to $P_i$ ($i=1,...,m-1$) and
3390: $$
3391: \angle( \t{\xi}_i, \t{\eta}_i)=\pi, i=1,...,m-1.
3392: $$
3393: Here $\t{\xi}_i, \t{\eta}_i, \t{\zeta}_i$
3394: are directions in $\Si_{\t{x}_i}(X)$ which correspond to the directions $\xi_i, \eta_i, \zeta_i$ under the
3395: congruences $T_i\to P_i$. Our goal is to produce a flat triangle $T_m\subset A_m\subset X$ so that the above
3396: properties still hold.
3397: 
3398: Since we have a simple chain $(-\xi_m, \eta_m)$ in
3399: $(S_{x_m}, W_{x_m})$, it follows from Proposition \ref{simplechain}
3400: that there exists a point $\tilde{\eta}_m\in \Si_{\t{x}_m}(X)$ so that
3401: $$
3402: d(\tilde{\eta}_m, \t{\xi}_m)=\pi, \quad d(\tilde{\eta}_m,
3403: \zeta_m)=d_{ref}(\eta_m, \zeta_m).
3404: $$
3405: Let $A_m$ denote an apartment in $X$ which contains $\ol{o \t{x}_m}$ and
3406: such that $\tilde{\eta}_m$ is tangent to $A_m$. Construct a geodesic segment
3407: $\ol{\t{x}_m \t{x}_{m+1}}\subset A_m\subset X$ whose metric length equals the one of $\ol{x_m x_{m+1}}$
3408: and whose initial direction is $\tilde{\eta}_m$. This defines a flat triangle
3409: $$
3410: T_m=[o, \t{x}_m, \t{x}_{m+1}]\subset A_m.
3411: $$
3412: It is clear from the construction that the triangle $T_m$ is congruent to $P_m$, in particular,
3413: $d_{ref}(o, x_3)=d_{ref}(o, y_3)$. Observe also that
3414: $$
3415: \ol{\t{x}_{m-1} \t{x}_m}\cup \ol{\t{x}_m \t{x}_{m+1}}$$
3416: is a geodesic segment (because $\angle(\tilde{\eta}_m, \t{\xi}_m)=\pi$). See Figure \ref{unfold.fig}.
3417: 
3418: Therefore, by induction we obtain a geodesic triangle $T=[o, x_1, \t{x}_n]\subset X$, which
3419: is triangulated (from $o$) into flat geodesic triangles $T_i$ which are congruent  to $P_i$'s.
3420: We claim that $f(T)=P$. For each $i$ the folding $f$ sends the
3421: triangle $T_i$ to $P_i$, according to Proposition \ref{uniqueness}.
3422: Therefore $f(T)=P$.
3423: 
3424: 
3425: 
3426: \begin{figure}[tbh]
3427: \centerline{\epsfxsize=5in \epsfbox{unfold.eps}} \caption{\sl
3428: Unfolding a broken triangle.} \label{unfold.fig}
3429: \end{figure}
3430: 
3431: 
3432: \medskip
3433: As in the proof of Theorem \ref{T2}, the argument has to be modified
3434: in case when $x_i=o$ for some $i$, $x_i=p(t_i)$. Then the vector
3435: $\mu=p'_-(t_i)$ belongs to the negative chamber $-\De$ and the
3436: vector $\la=p'_+(t_i)$ belongs to the positive chamber $\De$. Since
3437: $p$ is a billiard path, there exists $w\in W_{sph}$ which sends
3438: $\mu$ to $\la$. Now the chain and billiard conditions imply that
3439: $x_i$ is the  only break-point in $p$. Thus we can take
3440: $$
3441: T:= \ol{ox_1} \cup p([0,t_i]) \cup w^{-1} p([t_i, 1]) \cup
3442: w^{-1}(\ol{ox_n}).
3443: $$
3444: This degenerate geodesic triangle (it is contained in the geodesic
3445: through the points $x_1, w(x_n)$) folds to $P$ under the projection
3446: $\P_\De: A\to \De$. \qed
3447: 
3448: 
3449: 
3450: 
3451: \medskip
3452: The same argument as above proves the following generalization of
3453: Corollary \ref{chain-folded0}
3454: 
3455: 
3456: \begin{cor}
3457: \label{chain-folded} Suppose that $P$ is a  polygon in $\De$ which
3458: is the composition
3459: $$
3460: \ol{ox} \cup   p \cup  q \cup  \ol{yo}.
3461: $$
3462: Assume that paths $p, q$ satisfy the simple chain condition. Then
3463: there exists a polygon $\tilde{P}\subset X$ of the form
3464: $$
3465: \ol{ox} \cup \tilde{p} \cup  \tilde{q} \cup \ol{\t{y}o}
3466: $$
3467: so that $f(\tilde{P})=P$, $f(\tilde{p})=p, f(\tilde{q})=q$,
3468: $f(\t{y})=y$, and $\tilde{p}, \tilde{q}$ are geodesic paths.
3469: \end{cor}
3470: 
3471: 
3472: 
3473: 
3474: 
3475: 
3476:  We now use our analysis of the folded triangles
3477: (polygons) to relate them to the Littelmann triangles (polygons).
3478: 
3479: 
3480: \section{Littelmann polygons}
3481: \label{LSpolygons}
3482: 
3483: 
3484: \subsection{LS paths}
3485: \label{LSpaths}
3486: 
3487: 
3488:  Let $R$ be a root system on a Euclidean vector space
3489: $V$, $W=W_{sph}$ be the finite Coxeter group associated with $R$,
3490: let $W_{aff}$ denote the affine Coxeter group associated to $R$.
3491: This root system $R$ is actually the {\em coroot system} for the one
3492: considered by Littelmann in \cite{Littelmann2}. Accordingly, we will
3493: switch weights to coweights, etc. We pick a Weyl chamber $\Del$ for $W$, this
3494: determines the positive
3495: roots and the simple roots in $R$. We get the Euclidean Coxeter
3496: complex $(A,W_{aff})$, where $A$ is the affine space corresponding
3497: to $V$. Given $x\in A$ let $W_x$ denote the stabilizer of $x$ in $W_{aff}$.
3498: 
3499: 
3500: \medskip
3501: Suppose we are given a  vector $\la\in  \De\subset V$, a
3502: sequence of real numbers
3503: $$
3504: \ul{a}=( a_0=0< a_1 < ... < a_r=1),
3505: $$
3506: and a sequence of vectors in $W\la$
3507: $$
3508: \ul\nu= (\nu_1, ... , \nu_r), \hbox{~~so that~~} \nu_1 > ... > \nu_r
3509: $$
3510: with respect to the order in Definition \ref{order}.
3511: 
3512: 
3513: \begin{defn}
3514: The pair $(\ul\nu, \ul{a})$ is called a {\em real (billiard) path}
3515: of the $\De$-length $\la$.
3516: \end{defn}
3517: 
3518: 
3519: \begin{defn}
3520: (P. Littelmann \cite{Littelmann2}.) A real path of $\De$-length
3521: $\la$ is called {\em rational} if  $\la$ is a coweight and all
3522: numbers $a_i$ a rational.
3523: \end{defn}
3524: 
3525: 
3526: \begin{rem}
3527: Littelmann uses the notion {\em path of type $\la$} rather than of
3528: the $\De$-length $\la$.
3529: \end{rem}
3530: 
3531: 
3532: Set $a_i':= a_i -a_{i-1}$, $i=1, 2,...,r$. The data $(\ul\nu,
3533: \ul{a})$ determines a piecewise-linear path $p\in {\mathcal P}$ whose restriction
3534: to each interval $[a_{i-1}, a_i]$ is given by
3535: \begin{equation}
3536: \label{path}
3537:  p(t)= \sum_{k=1}^{i-1} a_k' \nu_k + (t-a_{i-1})\nu_i,
3538: t\in [a_{i-1}, a_i].
3539: \end{equation}
3540: 
3541: 
3542: 
3543: 
3544: Our interpretation of real and rational paths is the one of a broken
3545: (oriented) geodesic $L$ in $V$. Each oriented geodesic subsegment of
3546: $L$ is parallel to a positive multiple of an element of $W\la$, thus
3547: $L$ is a billiard path. The break points of the above path are the
3548: points
3549: $$
3550: x_1=a_1\nu_1, \ldots, x_i=x_{i-1}+ a_i' \nu_{i}, \ldots
3551: $$
3552: Since $\sum_i a_i'=1$, is clear that
3553: $$
3554: \Length (L)=\la,
3555: $$
3556: in the sense of the definition in section \ref{distances}. This
3557: justifies our usage of the name $\De$-length $\la$ in the above
3558: definitions, rather than Littelmann's notion of {\em type}.
3559: 
3560: 
3561: 
3562: 
3563: Observe that given a piecewise-linear path $p(t)\in {\mathcal P}$ (parameterized
3564: with the constant speed) one can recover the nonzero vectors
3565: $\nu_i\in V$ and the numbers $a_i$ and $a_i'$.
3566: 
3567: 
3568: 
3569: 
3570: 
3571: 
3572: 
3573: 
3574: 
3575: 
3576: 
3577: 
3578: \begin{defn}
3579: \label{LSdef} (P. Littelmann \cite{Littelmann2}.)
3580: A rational path
3581: $p(t)$ is called an {\em LS path}\footnote{a Lakshmibai-Seshadri
3582: path} if it satisfies further {\em integrality condition}:
3583: 
3584: 
3585: For each $i=1,...,s-1$ there exists an $a_i$-chain for the pair
3586: $(\nu_i, \nu_{i+1})$ (in the sense of Definition \ref{achain}).
3587: \end{defn}
3588: 
3589: 
3590: Observe that, since
3591: $$
3592: \eta \ge \tau \iff -\tau \ge -\eta,
3593: $$
3594: it follows that $p$ is an LS path if and only if $p^*$ is.
3595: 
3596: 
3597: 
3598: 
3599: \begin{thm}
3600: \label{locint}
3601:  (P.\ Littelmann \cite[Lemma 4.5]{Littelmann2}) Each LS path belongs to
3602: ${\mathcal P}_{\Z,loc}$.
3603: \end{thm}
3604: 
3605: 
3606: Our next goal is to give a more geometric interpretation of LS
3607: paths. Suppose that $p\in {\mathcal P}$ is a billiard path given by
3608: the equation (\ref{path}), with the vertices
3609: $$
3610: 0=x_0, x_1,...,x_r.
3611: $$
3612: At each vertex point $x_i, 0<i<r$,  we have unit tangent vectors
3613: $\xi_i, \mu_i$ which are tangent to the segments $\ol{x_i x_{i-1}},
3614: \ol{x_i  x_{i+1}}$. Note that at each vertex $x_i, 0<i<r$ we have
3615: the restricted and unrestricted spherical Coxeter complexes; the
3616: positive chamber $\De$ in $V$ determines positive chambers $\De_i$
3617: in the restricted spherical complexes $(S_{x_i}, W_{x_i})$.
3618: 
3619: 
3620: 
3621: 
3622: \begin{thm}
3623: \label{T3} A billiard path $p(t)$ of $\De$--length $\la\in P(R^\vee)$
3624: is an LS path if and only if it is a Hecke path which satisfies the
3625: {\em maximal chain condition} (cf. Definition \ref{D3}): At each
3626: vertex $x_i, 0<i<r$ there exists a $(S_{x_i}, W_{x_i},-\De_i)$-chain
3627: between $-\xi_i$ and $\mu_i$ and this chain is maximal as a
3628: $(S_{x_i}, W)$-chain.
3629: \end{thm}
3630: \proof Recall that given a nonzero vector $v\in V$, $\bar{v}$
3631: denotes its normalization $v/|v|$.
3632: 
3633: 
3634: 
3635: 
3636: 
3637: 
3638: \begin{figure}[tbh]
3639: \centerline{\epsfxsize=4.5in \epsfbox{LS.eps}} \caption{\sl
3640: Unbending a path.} \label{LS.fig}
3641: \end{figure}
3642: 
3643: 
3644: 
3645: 
3646: It is easy to see (and left to the reader) that if $p(t)$ is a
3647: satisfies the above chain condition and $\la=\Length(p)$ is a
3648: coweight, then all numbers $a_i$ are rational.
3649: 
3650: 
3651: Consider the first break point $x_1=x_0+a_1\nu_1$ of the broken
3652: geodesic path $p(t)$. Observe that
3653: $$
3654: \bar\nu_1=-\xi_1, \bar\nu_2=\mu_1\in S_{x_1}.
3655: $$
3656: According to Proposition \ref{eq}, existence of an $a_1$-chain for
3657: the pair $(\nu_1, \nu_{2})$ is equivalent to existence of an
3658: $(S_{x_1}, W_{x_1}, \De_1)$-chain, which is maximal in the
3659: unrestricted Coxeter complex,
3660: $$
3661: (\bar\nu_1=\eta_{1,0}, \eta_{1,1},...,\eta_{1,s_1}=\bar\nu_2).
3662: $$
3663: Thus the path $p$ satisfies the maximal chain  condition at the
3664: first break point $x_1$ if and only if it satisfies the {\em
3665: integrality condition} as in Definition \ref{LSdef}, at the point
3666: $x_1$.
3667: 
3668: 
3669: We now proceed to the next break point $x_2=x_1+ a_2' \nu_2$. We
3670: identify normalized vectors $\bar\nu_1, \bar\nu_2$ with unit vectors
3671: in $S_{x_1}$. Note that if $p(t)$ is an LS path of the $\De$-length
3672: $\la$, then there exists an element $w_1\in W_{x_1}$ which sends
3673: $\bar\nu_1$ to $\bar\nu_2$. The same is true if $p$ is a Hecke path.
3674: 
3675: 
3676: 
3677: 
3678: Set $x_2':= w_1^{-1}(x_2)$. Observe that, in both cases of an LS
3679: path and a Hecke path,
3680: $$
3681: \ol{x_0 x_1}\cup \ol{x_1 x_2'}
3682: $$
3683: is a geodesic segment $\ol{x_0 x_2'}$; the corresponding directed
3684: segment represents the vector $a_2 \nu_1$. Let $w_1'\in W_{sph}$
3685: denote the linear part of $w_1$. Set $x_2'':=w_1'(x_2')$. We
3686: translate the vectors $\bar\nu_2, \bar\nu_3$ to the unit tangent
3687: sphere $S_{x_2''}$.  The directed segment $\ora{x_0 x_2''}$
3688: represents the vector $a_2 \nu_2$. See Figure \ref{LS.fig}.
3689: 
3690: 
3691: We are now again in position to apply Proposition \ref{eq} with
3692: $a=a_2$: There exists a maximal chain
3693: $$
3694: (\bar\nu_2=\eta_{2,0}, \eta_{2,1},...,\eta_{2,s_2}=\bar\nu_3)
3695: $$
3696: if and only if there exists an $a_2$-chain for the pair
3697: $(\nu_2,\nu_3)$. The product $w_1\circ (w_1')^{-1}$ is a translation
3698: in $W_{aff}$ which carries $x_2''$ back to $x_2$. Therefore it
3699: induces an isomorphism of the restricted Coxeter spherical complexes
3700: $$
3701: (S_{x_2''}, W_{x_2''})\to (S_{x_2}, W_{x_2})
3702: $$
3703: which carries positive chamber to positive chamber. Hence this
3704: translation sends the chain $(\eta_{2,i})$ to a maximal chain in
3705: $(S_{x_2}, W_{x_2})$.
3706: 
3707: 
3708: We continue in this fashion: On the $i$-th step we ``unbend'' the
3709: broken geodesic
3710: $$
3711: \ol{x_0 x_1}\cup ... \cup\ol{x_{i-1} x_i}
3712: $$
3713: to a directed geodesic segment $\ora{x_0 x_i'}$ representing the
3714: vector $a_i\nu_1$, then apply an appropriate element $w_{i-1}'\in
3715: W_{sph}$ to transform segment $\ol{x_0 x_i'}$  to $\ol{x_0 x_i''}$;
3716: finally, appeal to Proposition \ref{eq} to establish equivalence
3717: between the maximal chain condition and the LS path axioms. \qed
3718: 
3719: 
3720: As a corollary of Theorem \ref{T3} we obtain
3721: 
3722: 
3723: \begin{cor}
3724: \label{C3} Let $T=[z,x,y]\subset X$ be a geodesic triangle and
3725: $f=Fold_{z,h,\De}$ be a folding into the Weyl chamber. Set
3726: $\be:=d_\De (x,y)$. Assume that $T'=f(T)$ is such that $f(x), f(y)$
3727: and all break-points of the broken geodesic $f(\ol{xy})$ are special
3728: vertices. Then $f(\ol{xy})$ is an LS path of the $\De$-length $k\be$.
3729: Here $k$ is the conformal factor of the dilation $h$.
3730: \end{cor}
3731: 
3732: 
3733: 
3734: 
3735: 
3736: 
3737: \subsection{Root operators}
3738: \label{operators}
3739: 
3740: 
3741: With each simple root $\al\in \Phi$, Littelmann
3742: \cite{Littelmann2} associates {\em raising and lowering} root
3743: operators $e_\al$ and $f_\al$ acting ${\mathcal P}$ as follows.
3744: 
3745: 
3746: Recall that given a path $p(t)$ and a root $\al$ we have the height function
3747: $h_\al(t):=\al(p(t))$. The number $m_\al$ is the minimal value of
3748: $h_\al$ on $[0,1]$.
3749: 
3750: 
3751: If $m_\al>-1$ then $e_\al$ is not defined on $p$. Otherwise let
3752: $t_1$ be the minimal $t$ for which $h_\al(t)=m_\al$ and let $t_0\in
3753: [0, t_1]$ be maximal such that $h_\al(t)\ge m_\al+1$ for all $t\in
3754: [0, t_0]$.
3755: 
3756: 
3757: The operator $e_\al$ will not change the path $p$ for $t\in [0,
3758: t_0]$ and, as far as $[t_1, 1]$ is concerned, the path $p|[t_1, 1]$
3759: will change only by a translation in $W_{aff}$ along the line
3760: $L_\al$ parallel to the vector $\al^\vee$. Thus it remains to
3761: describe the path $q=e_\al(p)$ on $[t_0, t_1]$. If $h_\al$ were not
3762: to have any local minima on $[t_0, t_1]$ then $q|_{[t_0, t_1]}$
3763: would be obtained by the reflection
3764: $$
3765: q|_{[t_0, t_1]}:= \tau_\al\circ p|_{[t_0, t_1]}
3766: $$
3767: and we would set
3768: $$
3769: q:= p|_{[0,t_0]} *\tau_\al\circ p|_{[t_0, t_1]} * p|_{[t_1,1]}.
3770: $$
3771: (Here we treat the paths resulting from the restriction of $p$ to
3772: subintervals of $[0,1]$ as elements of ${\mathcal P}$, according to
3773: the convention in section \ref{prelim}.)
3774: 
3775: 
3776: This is the definition of $e_\al$ of \cite{Littelmann1}, however the definition of
3777: $e_\al$ which we will need in this paper is the more refined
3778: one of \cite{Littelmann2}. Call a  subinterval $[s,u]\subset
3779: [t_0, t_1]$ a {\em spike} if it is a maximal interval satisfying
3780: $$
3781: h_\al(s)=h_\al(u) =\min( h_\al \restr [s,u]).
3782: $$
3783:  Thus $h_\al\restr [t_0,t_1]$ is decreasing on the complement to the
3784: union of {\em spikes}. The restriction of $q$ to each {\em spike} is
3785: obtained from $p$ by a translation along $L_\al$. The restriction to
3786: each subinterval disjoint from a {\em spike} is obtained by a
3787: reflection. To be more precise, subdivide the interval $[t_0,t_1]$
3788: into
3789: $$
3790: [t_0, s_1]\cup [s_1,s_2]\cup ... \cup [s_k, t_1],
3791: $$
3792: where the {\em spike} and {\em non-spike} intervals alternate. Observe
3793: that $[t_0, s_1], [s_k, t_1]$ are not {\em spikes}. Then
3794: $$
3795: q:= p|_{[0,t_0]} *\tau_\al(p|_{[t_0, s_1]}) * p|_{[s_1,s_2]}*...
3796: *\tau_\al( p|_{[s_k, t_1]}) * p|_{[t_1,1]}.
3797: $$
3798: 
3799: 
3800: 
3801: 
3802: Note that the operator $e_\al$ changes the geometry of the path $p$
3803: by an isometry near every point $p(t)$ which is neither a point of
3804: local minimum for $h_\al$ nor is a point where $h_\al(t)=m_\al-1$.
3805: Otherwise the local change is done by a ``bending''  with respect to
3806: a hyperplane parallel to $H_\al$. These hyperplanes are not
3807: necessarily walls of $W_{aff}$. However, if all local minimal values
3808: of $h_\al$ belong to $\Z$, these hyperplanes are indeed walls and we
3809: obtain:
3810: 
3811: 
3812: For each path $p\in {\mathcal P}_{\Z,loc}$, for each simple root
3813: $\al$, the path $q=e_\al(p)$ satisfies the following: The interval
3814: $[0,1]$ can be subdivided into subintervals $[s_i, s_{i+1}]$ such
3815: that the restriction $q|[s_i, s_{i+1}]$ is obtained from the
3816: restriction of $p$ by post-composition with an element of $W_{aff}$.
3817: 
3818: 
3819: \medskip
3820: The lowering operators $f_\al$ are defined analogously to the
3821: raising operators, we refer the reader to \cite{Littelmann2} for the
3822: precise definition. (See however Property 1 below.) At this stage we
3823: note only that $f_\al$ is undefined on $p$ iff $m_\al> h_\al(1)-1$.
3824: Let ${\mathcal E}$ be the semigroup generated by $e_\al$'s,
3825: ${\mathcal F}$ be the semigroup generated by $f_\al$'s and
3826: ${\mathcal A}$ be the semigroup generated by all root operators. The
3827: semigroups contain the identity operator by default. For each
3828: $\phi\in {\mathcal A}$ let $Dom(\phi)$ denote the domain of $\phi$.
3829: 
3830: 
3831: 
3832: 
3833: 
3834: 
3835: \begin{rem}
3836: In fact, Littelmann extends the operators $f_\al, e_\al$ to the
3837: entire ${\mathcal P}$ by declaring $f_\al(p)={\mathbf 0}$ for all
3838: $p$ for which $f_\al$ is undefined. However we will not need this
3839: extension in the present paper.
3840: \end{rem}
3841: 
3842: 
3843: Below we list certain properties of the root operators. Most of them
3844: are either clear from the definition or are proven in
3845: \cite{Littelmann2}. Most proofs that we present are slight
3846: modifications of the arguments in \cite{Littelmann2}.
3847: 
3848: 
3849: \medskip
3850: {\bf Property 1.} (P. Littelmann, \cite[Lemma 2.1
3851:  (b, e)]{Littelmann2}.)
3852: $$
3853: e_\al\circ f_\al(p)= p, \hbox{~~if~~} p\in Dom(f_\al),
3854: $$
3855: $$
3856: f_\al\circ e_\al(p)= p, \hbox{~~if~~} p\in Dom(e_\al),
3857: $$
3858: $$
3859: e_\al(p^*)= (f_\al(p))^*, \quad (e_\al(p))^*= f_\al(p^*),
3860: $$
3861: the latter could be taken as the definition of $f_\al$.
3862: 
3863: 
3864: 
3865: 
3866: \medskip
3867: {\bf Property 2.} \cite[Lemma 2.1]{Littelmann2}. For each
3868: $p\in Dom(e_\al)\cap {\mathcal P}$,
3869: $$
3870: m_\al(e_\al(p))= m_\al(p)+1,
3871: $$
3872: $$
3873: p\in Dom(e^N_\al) \iff N < |m_\al|.
3874: $$
3875: 
3876: 
3877: 
3878: 
3879: 
3880: 
3881: \medskip
3882: {\bf Property 3.} Suppose that $p$ is a path in ${\mathcal P}_\Z$
3883: which does not belong to the domain of any $e_\al, \al\in \Phi$.
3884: Then $p$ is contained in $\Del$. Indeed, for each simple root $\al$
3885: we have to have $m_\al(p)>-1$. Since $p\in {\mathcal P}_\Z$,
3886: $m_\al(p)=0$. Thus $p\in {\mathcal P}^+$.
3887: 
3888: 
3889: \medskip
3890: {\bf Property 4.}  \cite[Proposition 3.1 (a, b)]{Littelmann2}.
3891: For each $\al \in \Phi$, $Dom(f_\al)\cap {\mathcal P}_\Z$ is open
3892: and $f_\al|{\mathcal P}_\Z$ is continuous.
3893: 
3894: 
3895: \medskip
3896: {\bf Property 5.} \cite[\S 7, Corollary 1 (a)]{Littelmann2}.
3897: Let $p\in {\mathcal P}^+$ and $\phi$ be a composition of lowering
3898: operators defined on $p$. Then $\phi(p)\in {\mathcal P}_\Z$ .
3899: 
3900: 
3901: \medskip
3902: {\bf Property 6.} Combining Properties 4 and 5 we conclude that for
3903: each ${\mathbf f}\in {\mathcal F}$, $Dom({\mathbf f})\cap {\mathcal
3904: P}^+$ is open and ${\mathbf f}|{\mathcal P}^+$ is continuous.
3905: 
3906: 
3907: \medskip
3908: {\bf Property 7.} \cite[Corollary 3, Page 512]{Littelmann2}.
3909: $p$ is an LS path of the $\De$-length $\la$ if and only if there
3910: exists ${\mathbf f}\in {\mathcal F}$ such that
3911: $$
3912: p={\mathbf f}(\pi_\la)
3913: $$
3914: 
3915: 
3916: \medskip
3917: {\bf Property 8.} \cite[Corollary 2(a), page
3918: 512]{Littelmann2}. The set of LS paths of the given $\De$-length is
3919: stable under ${\mathcal A}$.
3920: 
3921: 
3922: \medskip
3923: {\bf Property 9.} Suppose that $p\in {\mathcal P}$, $t\in [0,1]$,
3924: $\al\in \Phi$ and $x:=p(t)$ satisfy
3925: $$
3926: \al(x)\in \Z,  \quad p'_-(t)\gtrsim_{W_x} p'_+(t).
3927: $$
3928: Then the path $q=e_\al(p)$ also satisfies
3929: $$
3930: q'_-(t)\gtrsim_{W_y} q'_+(t),
3931: $$
3932: for $y=q(t)$. \proof If $h_\al(t)\ne m_\al, m_\al-1$, the germs of
3933: the paths $p$ and $q$ at $t$ differ by a translation. Thus the
3934: conclusion trivially holds in this case. The same argument applies
3935: if $h_\al(t)=m_\al-1$ and
3936: $$
3937: \al(p_-'(t)) \ge 0, \al(p_+'(t))\le 0.
3938: $$
3939: The nontrivial cases are:
3940: 
3941: 
3942: 1. $h_\al(t)=m_\al-1$, $\al(p'_-(t))\le 0$, $\al(p'_+(t))\le 0$. In
3943: this case the assertion follows from Lemma \ref{easy} with
3944: $\nu=p'_-(t), \mu=p'_+(t)$.
3945: 
3946: 
3947: 2. $h_\al(t)=m_\al$, $\al(p'_-(t))\le 0$, $\al(p'_+(t))\ge 0$. In
3948: this case the assertion follows from Lemma \ref{hard} with
3949: $\nu=p'_-(t), \mu=p'_+(t)$. \qed
3950: 
3951: 
3952: 
3953: 
3954: \medskip
3955: {\bf Property 10.}  Suppose that $p=p_1*p_2$ where $p_1\in {\mathcal
3956: P}^+$. Then for each $e\in {\mathcal E}$ defined on $p$ we have
3957: $$
3958: e(p)= p_1* e(p_2).
3959: $$
3960: \proof It suffices to prove this for $e=e_\al$, $\al\in \Phi$. In
3961: the latter case it follows directly from the definition of the
3962: operator $e_\al$. \qed
3963: 
3964: 
3965: 
3966: 
3967: The next property is again clear from the definition:
3968: 
3969: 
3970: \medskip
3971: {\bf Property 11.} Suppose that
3972: $$
3973: {\mathbf e}= e_{\be_m}\circ ... \circ e_{\be_1}
3974: $$
3975: where $\be_i\in \Phi$, $p\in Dom({\mathbf e})$. Set
3976: $$
3977: p_i:= e_{\be_i}\circ ...\circ e_{\be_1}(p), i=1,...,m.
3978: $$
3979: Then for each $T\in [0, 1]$, the sequence of vectors
3980: $$
3981: (p'_+(T), (p_1)_+'(T),..., (p_m)_+'(T)),
3982: $$
3983: after deleting equal members, forms a chain.
3984: 
3985: 
3986: \begin{lem}
3987: \label{finiteness} Given a path $p$ there are only finitely many
3988: operators ${\mathbf e}\in {\mathcal E}$ which are defined on $p$.
3989: \end{lem}
3990: \proof Break the path $p$ as the concatenation
3991: $$
3992: p_1*...*p_s.
3993: $$
3994: of geodesic paths each of which is contained in a single alcove and
3995: let $T_i\in [0, 1]$ be such that $p(T_i)=p_i(1/2)$; set $T_0:=0$.
3996: Then for each $e_\al\in {\mathcal E}$ defined on $p$ there exists
3997: $i$ such that the derivatives of ${e}_\al(p)$ and $p$ at $T_i$ are
3998: not the same. Moreover,
3999: $$
4000: q={e}_\al(p)= q_1*...*q_s,$$ where each $q_i$ is a geodesic path
4001: contained in an alcove. Consider the vector
4002: $$
4003: L(p):= (\ell( p'(0)), \ell(p'(T_1)),..., \ell(p'(T_s))) \in (\N\cup
4004: \{0\}) ^{s+1},
4005: $$
4006: where $\N^{s+1}$ is given the lexicographic order and $\ell$ is the length
4007: function on $W_{sph}$-orbits induced from the word metric on
4008: $W_{sph}$ as in Proposition \ref{wordlength}. Then, by combining Proposition
4009: \ref{wordlength} and Property 11 above, for each $\al\in \Phi$,
4010: $$
4011: L(e_\al(p))< L(p).
4012: $$
4013: Lemma follows. \qed
4014: 
4015: 
4016: 
4017: 
4018: 
4019: 
4020: 
4021: 
4022: 
4023: 
4024: \subsection{Generalized LS paths}
4025: \label{GLS}
4026: 
4027: 
4028: 
4029: 
4030:  In this paper we will need two generalizations of the
4031: concept of an LS path; the first one will be needed for the proof of
4032: the saturation theorem (section \ref{saturationsection}), the second
4033: will be used in section \ref{unfoldingLS} for the proof of the
4034: unfolding theorem. Although we will use the name {\em generalized LS
4035: path} for both generalizations, it will be clear from the context
4036: which generalization is being referred to.
4037: 
4038: 
4039: \medskip
4040: {\bf The first generalization, ${\mathcal L}{\mathcal S}_1$}.
4041: \footnote{This generalization of LS paths will be used in the proof of the saturation theorem.}
4042: 
4043: \medskip
4044: Suppose we are given a collection of LS paths $p_i$ of the
4045: $\De$-length $\la_i\in \De\cap P(R^\vee)$, $i=0,...,m$. We will use
4046: the notation
4047: $$
4048: \ul{\la}=(\la_0,...,\la_m)
4049: $$
4050: and
4051: $$
4052: \la:= \sum_{i=0}^m \la_i.
4053: $$
4054: 
4055: 
4056: \begin{rem}
4057: Actually, for our main application it will suffice to consider
4058: $\la_i$'s which are multiples of the fundamental coweights
4059: $\varpi_i$. Therefore such paths are automatically {\em
4060: generalized Hecke paths} as defined in Definition \ref{gH}.
4061: \end{rem}
4062: 
4063: 
4064: 
4065: 
4066: \begin{defn}
4067: \label{LS1} The concatenation
4068: $$
4069: p= p_0*p_1*...*p_m
4070: $$
4071: will be called a {\em generalized LS path} with $\length
4072: (p)=\ul{\la}$, if for each $i=0,...,m-1$
4073: $$
4074: p'_i(1) \gtrsim  p'_{i+1}(0).
4075: $$
4076: The set of such generalized LS paths will be denoted ${\mathcal
4077: L}{\mathcal S}_1$.
4078: \end{defn}
4079: 
4080: 
4081: This definition is a very special case of the one used by Littelmann
4082: in \cite{Littelmann3} under the name of a {\em locally integral
4083: concatenation}.
4084: 
4085: 
4086: Recall that according to the definition of $\De$-length,
4087: $$
4088: \la= \Length(p).
4089: $$
4090: Observe that each LS path $p$ satisfies the above definition, since
4091: for each $i$
4092: $$
4093: p'_i(1) \ge p'_{i+1}(0).
4094: $$
4095: 
4096: 
4097: \begin{ex}
4098: Suppose that $u, v$ are dominant coweights. Then $p=\pi_u*\pi_v$ is
4099: a generalized LS path.
4100: \end{ex}
4101: 
4102: 
4103: 
4104: 
4105: 
4106: \medskip
4107: {\bf The second generalization ${\mathcal L}{\mathcal S}_2$}.
4108: \footnote{This notion of generalized LS path will be used only to unfold Hecke paths.}
4109: 
4110: 
4111: \medskip
4112: Suppose that $p_1, p_2\in {\mathcal P}$ appear as
4113: $$
4114: p_1= \tilde{p}_1|_{[0,a]}, \quad 0<a<1,
4115: $$
4116: $$
4117: p_2= \tilde{p}_2|_{[b,1]}, \quad 0<b<1,
4118: $$
4119: where $\tilde{p}_1, \tilde{p}_2$ are LS paths, $a, b\in \Q$. (See
4120: section \ref{paths} for the definition of $\tilde{p}_1|_{[0,a]}$
4121: and $\tilde{p}_2|_{[b,1]}$.) Define the path $p:= p_1*p_2$. Assume
4122: that
4123: $$
4124: p_1'(1)\rhd p_2'(0);
4125: $$
4126: in other words, if $t$ is such that $p(t)= p_1(1)$ then
4127: $$
4128: p'_-(t) \rhd p_+(t).
4129: $$
4130: 
4131: \begin{defn}
4132: The concatenation $p$ will be called a {\em generalized LS path}
4133: if the {\em concatenation point} $p_1(1)$ is a regular
4134: point\footnote{I.e. it does not belong to any wall.} of
4135: $(A,W_{aff})$ and $p(1)\in P(R^\vee)$.
4136: 
4137: 
4138: The set of such generalized LS paths will be denoted ${\mathcal
4139: L}{\mathcal S}_2$.
4140: \end{defn}
4141: 
4142: 
4143: \begin{ex}
4144: \label{e2} Suppose that $u\in\De, v\in V$ are such that $u, v \in
4145: P(R^\vee)\otimes \Q$, $u+v\in P(R^\vee)\cap \De$ and the head of the
4146: vector $u$ is a regular point in $(A,W_{aff})$. Then $p=\pi_u*
4147: \pi_v\in {\mathcal L}{\mathcal S}_2$.
4148: \end{ex}
4149: 
4150: 
4151: This definition is again a very special case of the one given by
4152: Littelmann in \cite[5.3]{Littelmann2}. Littelmann does not assume
4153: that $p_1(1)$ is regular, but instead imposes certain chain
4154: conditions at this point.
4155: 
4156: 
4157: \medskip
4158: {\bf Properties of generalized LS paths:}
4159: 
4160: 
4161: \medskip
4162: {\bf Property 0.} If $p\in {\mathcal L}{\mathcal S}_1$ then $p^*$ is
4163: also in ${\mathcal L}{\mathcal S}_1$.
4164: 
4165: 
4166: \proof Represent $p$ as a concatenation $p_1*...*p_m$ of LS paths as
4167: in the Definition \ref{LS1}. Then
4168: $$
4169: p^*=(p_m^*)*...*(p_1^*),
4170: $$
4171: where each path $p_i^*$ is again an LS path. Now the assertion
4172: follows from Lemma \ref{iff}. \qed
4173: 
4174: 
4175: \medskip
4176: {\bf Property 1.} ${\mathcal L}{\mathcal S}_2$ is stable under the
4177: root operators  \cite[Lemma 5.6, 2-nd assertion]{Littelmann2}.
4178: In particular, suppose that $p$ is as in Example \ref{e2}. Then for
4179: each ${\mathbf f}\in {\mathcal F}$  (defined on $p$), ${\mathbf
4180: f}(p)$ is a generalized LS path of the $\De$-length $\la=u+v$.
4181: 
4182: 
4183: {\bf Property 2.}
4184:  ${\mathcal L}{\mathcal S}_1$ is stable
4185:  under the root operators.
4186: 
4187: 
4188: \proof Suppose that $p=p_1*...*p_m$ is a concatenation of LS paths
4189: as above and $e_\al$ is a raising operator. In particular, for each
4190: $i$ we have a vector $u_i$ so that
4191: $$
4192: p'_i(1) \ge u_i \sim p'_{i+1}(0).
4193: $$
4194: For each $i$, $e_\al(p_i)$ is again a Littelmann path. Therefore
4195: $e_\al(p)$ is a concatenation of LS paths $q_1*...*q_m$. We have to
4196: verify that for each $i$ there is a vector $v_i\in V$ so that
4197: $$
4198: q'_i(1) \ge v_i \sim q'_{i+1}(0).
4199: $$
4200: This however follows from the Property 9 in the previous section. To
4201: check that ${\mathcal L}{\mathcal S}_1$ is preserved by $f_\al$ we
4202: use that $q\in {\mathcal L}{\mathcal S}_1\iff q^* \in {\mathcal
4203: L}{\mathcal S}_1$ and
4204: $$
4205: f_\al(p)= (e_\al(p^*))^*.  \qed
4206: $$
4207: 
4208: 
4209: 
4210: 
4211: \medskip
4212: {\bf Property 3.} ${\mathcal L}{\mathcal S}_1$ and ${\mathcal
4213: L}{\mathcal S}_2$ are contained in ${\mathcal P}_{\Z,loc}$. For
4214: ${\mathcal L}{\mathcal S}_1$ it is immediate since, by Theorem
4215: \ref{locint}, the set of LS
4216: paths is contained in ${\mathcal P}_{\Z,loc}$.
4217: For ${\mathcal L}{\mathcal S}_2$ it is a special case
4218: of \cite[Lemma 5.5]{Littelmann2}.
4219: 
4220: 
4221: \medskip
4222: {\bf Property 4.} Suppose that $p\in {\mathcal L}{\mathcal S}_1$.
4223: Then there exists an element ${\mathbf e}\in {\mathcal E}$
4224: defined on $p$ such that $q={\mathbf e}(p)\in {\mathcal P}^+$.
4225: 
4226: 
4227: \proof If $m_{\al_1}(p)\le -1$ then we apply a power
4228: $e_{\al_1}^{k_1}$ to $p$ so that $q_1:=e_{\al_1}^{k_1}(p)$ satisfies
4229: $m_{\al_1}(q_1)>-1$. However, since $e_{\al_1}^{k_1}(p)\in {\mathcal
4230: L}{\mathcal S}_1\subset {\mathcal P}_{\Z,loc}$, it follows that
4231: $q_1\in {\mathcal P}_{\Z,loc}$ and so $m_{\al_1}(q_1)=0$. We then
4232: apply a power of $e_{\al_2}$ to $q_1$, etc. According to Lemma
4233: \ref{finiteness}, this process must terminate. Therefore, in the end
4234: we obtain a path
4235: $$
4236: {\mathbf e}(p)=q
4237: $$
4238: which does not belong to the domain of any raising operator. Since
4239: $q\in {\mathcal P}_{\Z,loc}$ it follows that $q$ is entirely
4240: contained in $\De$. \qed
4241: 
4242: 
4243: Recall  \cite[Chapter VI, section 10]{Bourbaki} that if a root system
4244: $R$ spans $V$ then each dominant coweight $\la\in \De$ is a
4245: positive integral combination
4246: $$
4247: \la= \sum_{i=1}^l n_i \varpi_i,
4248: $$
4249: where $\varpi_i$ are fundamental coweights. This assertion (as it stands) is false without the above assumption
4250: on $V$. In the general case we  have
4251: $$
4252: \la= \la'+ \sum_{i=1}^l n_i \varpi_i,
4253: $$
4254: where $\la'\in V'$, $n_i\in \N\cup \{0\}$.
4255: As an alternative the reader can restrict the discussion to semisimple groups only,
4256: when $V'=0$.
4257: 
4258: 
4259: \begin{conv}
4260: \label{con} From now on we will be assuming that in Definition
4261: \ref{LS1}
4262: $$
4263: \la_j=k_j \varpi_j, \quad k_j\in \N,
4264: $$
4265: for each $j=1,...,m$, where $\varpi_j$ is the $j$-th fundamental coweight, and
4266: $$
4267: \la_0\in V'.
4268: $$
4269: Then the subpath $p_0$ is necessarily geodesic.
4270: \end{conv}
4271: 
4272: 
4273: \begin{lem}
4274: \label{positiv} Suppose that $p\in {\mathcal L}{\mathcal S}_1 \cap
4275: {\mathcal P}^+$ is a generalized LS path with $\length(p)=\ul{\la}$.
4276: Then
4277: $$
4278: p=\pi_{\la_0}*...*\pi_{\la_m}.
4279: $$
4280: \end{lem}
4281: \proof Represent $p$ as the concatenation of maximal LS subpaths,
4282: $p=p_0*p_1...*p_m$. The geodesic subpath $p_0$ clearly equals $\pi_{\la_0}$.
4283: Since $p_1$ is an LS path and $p_1'(0)\in \De$, we
4284: see that $p_1$ is a geodesic path (see Corollary \ref{end}) which
4285: therefore equals $\pi_{\la_1}$. Moreover, because
4286: $$
4287: p_1'(1)\ge u_1 \sim p'_2(0),
4288: $$
4289: it follows that $u_1=p_1'(1)$ and thus $p_1'(1)=\la_1\sim p_2'(0)$.
4290: Let $x_1:= p_1(1)$; this point lies on the boundary face of $\De$
4291: which does not contain $\la_2$. Note that the vector $p_2'(0)$,
4292: regarded as an element of $T_{x_1}(A)$, points inside the Weyl
4293: chamber $\De$ (for otherwise $p$ is not contained in $\De$). On the
4294: other hand, since $\la_1\sim p_2'(0)$, the vector $p_2'(0)$ belongs
4295: either to $\De$ or to the Weyl chamber
4296: $$
4297: \tau_{\be_2}(\De)
4298: $$
4299: adjacent to $\De$. Since $p\in {\mathcal P}^+$, it is clear that
4300: $p_2'(0)\in \De$. Thus $p_2$ is the geodesic path $\pi_{\la_2}$.
4301: Continuing in this fashion we conclude that
4302: $$
4303: p=\pi_{\la_0}*\pi_{\la_1}*...*\pi_{\la_m}. \qed
4304: $$
4305: 
4306: 
4307: \begin{thm}
4308: \label{pathlow}
4309:  Suppose that $p$ is a generalized LS path in the sense of ${\mathcal L}{\mathcal S}_1$
4310:  with $\length(p)=\ul{\la}$
4311: (satisfying convention \ref{con}). Then there exists ${\mathbf f}\in
4312: {\mathcal F}$ such that
4313: $$
4314: p={\mathbf f}(\pi_{\la_0}*...*\pi_{\la_m})
4315: $$
4316: \end{thm}
4317: \proof If $p\in {\mathcal P}^+$ then we are done. Otherwise, by
4318: combining Lemma \ref{positiv} with the Property 4, we find an
4319: ${\mathbf e}\in {\mathcal E}$,
4320: $$
4321: {\mathbf e}= e_{\al_1}^{k_1}\circ ... e_{\al_n}^{k_n}
4322: $$
4323:  such that $p\in Dom({\mathbf e})$ and ${\mathbf e}(p)=q\in
4324: {\mathcal P}^+$. Therefore
4325: $$
4326: q=\pi_{\la_0}*\pi_{\la_1}*...*\pi_{\la_m}
4327: $$
4328: and thus the composition
4329: $$
4330: {\mathbf f}= f_{\al_n}^{k_n}\circ ... f_{\al_1}^{k_1}
4331: $$
4332: satisfies $p={\mathbf f} {\mathbf e}(p)={\mathbf
4333: f}(\pi_{\la_0}*\pi_{\la_1}*...\pi_{\la_m})$. \qed
4334: 
4335: 
4336: 
4337: 
4338: \subsection{Path model for the representation theory of Lie groups}
4339: \label{LSrep}
4340: 
4341: 
4342: Suppose that
4343: $$p(t)\in {\mathcal L}{\mathcal S}_1 \cup {\mathcal L}{\mathcal S}_2$$
4344: is a generalized LS path with $\Length(p)=\be$ and
4345: $$\length(p)=\ul{\la}.$$
4346: Suppose that $\al\in P(R^\vee)$ is such that $\al+p(t)$ is contained
4347: in $\De$. Then $\al$ and $p$ define a polygon
4348: $$
4349: P:=\ol{o y_0} \cup (p+\al)\cup \ol{y_n o} \subset \De$$
4350:  where $\al=\ov{o y_0}$, $y_n=\al+p(1)$. Let $\ga$ denote the vector
4351: $\ov{o y_{n}}$; then $\ga$ is also a dominant coweight. Recall that
4352: the contragredient dominant coweight $\ga^*\in \De$ is obtained by
4353: projecting the vector $-\ga$ to the Weyl chamber $\De$ by the
4354: projection $\P: V\to \De$.
4355: 
4356: 
4357: \begin{defn}
4358: 1. A polygon $P$ above will be called a {\em (broken) Littelmann
4359: polygon} with the $\De$-side lengths $\al, \be, \ga^*$.
4360: 
4361: 
4362: 2. If $p(t)$ is an LS path then $P$ will be called a {\em (broken)
4363: Littelmann triangle} with the $\De$-side lengths $\al, \be, \ga^*$.
4364: \end{defn}
4365: 
4366: 
4367: Pick a lattice $L$ such that
4368: $$
4369: Q(R^\vee)\subset L\subset P(R^\vee).
4370: $$
4371: Then there exists a unique connected semisimple complex Lie group
4372: $G^\vee$  with the root system $R^\vee$ and the character lattice
4373: $L$ of the maximal torus $T^\vee\subset G^\vee$.
4374:  Recall that irreducible representations $V$ of $G^\vee$ are
4375: parameterized by their dominant weights, $V=V_\la$, $\la\in \De\cap
4376: L$.
4377: 
4378: 
4379: Pick a path $q\in {\mathcal P}^+$ such that $q(1)=\be$. Then,
4380: according to \cite[Decomposition formula, Page 500]{Littelmann2} we
4381: have
4382: 
4383: 
4384: \begin{thm}
4385: \label{Lit2} The tensor product $V_\al\otimes
4386: V_\be$  contains $V_\ga$ as a subrepresentation if and only if there
4387: exists a path $p\in {\mathcal F}(q)$ such that $\pi_{\al}*p\in
4388: {\mathcal P}^+$ and $\pi_{\al}*p(1)=\ga$.
4389: \end{thm}
4390: 
4391: 
4392: \begin{rem}
4393: Littelmann works with simply-connected group $G^\vee$ and weights
4394: $\al, \be,\ga$ in $P(R^\vee)$. The statement for
4395: non-simply-connected groups trivially follows from the
4396: simply-connected case.
4397: \end{rem}
4398: 
4399: 
4400: 
4401: 
4402: In particular, since $p$ is an LS paths of the $\De$-length $\be$ if
4403: and only if $p\in {\mathcal F}(\pi_\be)$, it follows that
4404: 
4405: 
4406: \begin{thm}
4407: \cite{Littelmann1, Littelmann2}. \label{Lit} The tensor
4408: product $V_\al\otimes V_\be$  contains $V_\ga$ as a
4409: subrepresentation if and only if there exists a (broken) Littelmann
4410: triangle in $\Del\subset V$, with the $\De$-side-lengths $\al, \be,
4411: \ga^*$.
4412: 
4413: 
4414: In other words, $V_\ga\subset V_\al\otimes V_\be$ if and only if
4415: there exists an LS path $p$ of $\De$-length $\be$ such that
4416: 
4417: 
4418: 1. $\pi_\al *p\in {\mathcal P}^+$.
4419: 
4420: 
4421: 2. $p(1)+\al=\ga$.
4422: \end{thm}
4423: 
4424: 
4425: 
4426: 
4427: 
4428: 
4429: We will apply Theorem \ref{Lit2} as follows.
4430: Represent the vector $\be$ as the integer linear combination of
4431: fundamental coweights
4432: $$
4433: \be =\sum_{i=1}^n k_i \varpi_i.
4434: $$
4435: We reorder the fundamental coweights so that $k_i>0$ for all
4436: $i=1,...,m$ and $k_i=0, i\ge m+1$. Set $\la_i:= k_i \varpi_i, 1\le
4437: i\le m$ and let $\ul{\la}= (\la_1,...,\la_m)$. Therefore the path
4438: $$
4439: \pi_{\ul\la}:= \pi_{\la_1}*... \pi_{\la_m}
4440: $$
4441: belongs to ${\mathcal L}{\mathcal S}_1$ and $\pi_{\ul\la}(1)=\be$.
4442: Then
4443: 
4444: 
4445: 
4446: 
4447: \begin{cor}
4448: The tensor product $V_\al\otimes V_\be$ contains $V_\ga$ as a
4449: subrepresentation
4450:  if and only if there exists a generalized LS path $p$ so that
4451: 
4452: 
4453: 1. $\length(p)=\ul{\la}$.
4454: 
4455: 
4456: 2. $\pi_{\al}*p\in {\mathcal P}^+$.
4457: 
4458: 
4459: 3. $\pi_{\al}*p(1)=\ga$.
4460: \end{cor}
4461: \proof Set
4462: $$
4463: q=\pi_{\ul{\la}}.$$
4464:  According to Theorems \ref{pathlow} and Property 2 of
4465: generalized LS paths (section \ref{GLS}), $p\in {\mathcal P}$ is a
4466: generalized LS path with $\length(p)=\ul\la$ if and only if $p\in
4467: {\mathcal F}(q)$. Now the assertion follows from Theorem \ref{Lit2}.
4468: \qed
4469: 
4470: 
4471: Combining Corollary \ref{folded-chain}, Theorem \ref{T3} and Theorem
4472: \ref{Lit} we obtain
4473: 
4474: 
4475: \begin{cor}
4476: Suppose that $X$ is a thick Euclidean building modeled on the
4477: Coxeter complex $(A,W_{aff})$. Let $\al, \be, \ga^*\in L$ be
4478: dominant coweights. Suppose that $a\subset A$ is an alcove
4479: containing a special vertex $o$, $T=[o, x, y]\subset X$ is a
4480: geodesic triangle with the special vertices  and the $\De$-side
4481: lengths $\al, \be, \ga^*$. Assume also that the broken side
4482: $Fold_{a,A}(\ol{x y})$ of the folded triangle
4483: $$
4484: Fold_{a,A}(T)
4485: $$
4486: has breaks only at the special vertices of $A$. Then
4487: 
4488: 
4489: 1. The folded triangle $T'=Fold_{o,id,\De}(T)\subset \De$ is a
4490: Littelmann triangle.
4491: 
4492: 
4493: 2. $V_\ga\subset V_\al\otimes V_\be$.
4494: \end{cor}
4495: \proof Indeed, according to Corollary \ref{folded-chain}, the folded
4496: triangle $T'$ satisfies the chain condition. Each break point $x_i$
4497: on the broken side of $T'$ is
4498: 
4499: 
4500: 1. Either a special vertex, in which case it satisfies maximal chain
4501: condition by Remark \ref{R1}, or
4502: 
4503: 
4504: 2. $Fold_{a,A}(\ol{xy})$ is geodesic at the point corresponding to
4505: $x_i$, so the chain at $x_i$ can be chosen to be maximal by
4506: Proposition \ref{chaincondition1}.
4507: 
4508: 
4509: Hence Theorem \ref{T3} implies that $T'$ is a Littelmann triangle.
4510: The second assertion now follows from Theorem \ref{Lit}. \qed
4511: 
4512: 
4513: \medskip
4514: Of course, the assumption that the break points occur only at the
4515: special vertices is very restrictive. In section
4516: \ref{saturationsection} we will get rid of this assumption at the
4517: expense of dilation of the side-lengths.
4518: 
4519: 
4520: \section{Unfolding}
4521: \label{last}
4522: 
4523: 
4524: 
4525: 
4526: The goal of this section is to establish an intrinsic
4527: characterization of folded triangles as the broken billiard
4528: triangles satisfying the chain condition. We first prove this
4529: characterization for Littelmann triangles and then, using this, give
4530: a general proof.
4531: 
4532: 
4533: 
4534: 
4535: Throughout this section we assume that $X$ is a thick locally
4536: compact Euclidean building modeled on the Coxeter complex $(A,W)$,
4537: $\Del\subset A$ is a Weyl chamber with tip $o$. Let $g: X\to
4538: \De$ denote the folding $Fold_{\De}$.
4539: 
4540: 
4541: 
4542: Let $T\subset \Del$ be a  billiard triangle which is the
4543: composition
4544: $$
4545: T=\ov{ox}\cup r \cup \ol{y o},
4546: $$
4547: where $r(t)=p(t)+\al$, $\al=\ov{ox}$ and $p\in {\mathcal P}$ is a Hecke path. Thus $T$
4548: has the geodesic sides $\ol{ox}, \ol{oy}$ and the broken side
4549: $r$. We set $\ga:= \ov{o y}$ and let $\be\in \De$ denote
4550: $\De$-length of the path $p$.
4551: 
4552: 
4553: 
4554: 
4555: \subsection{Unfolding Littelmann triangles}
4556: \label{unfoldingLS}
4557: 
4558: 
4559: \begin{thm}
4560: \label{LSfolded}
4561:  Suppose that, in addition, $T$ is a {\em Littelmann triangle},
4562: i.e. $\al, \ga\in L\subset P(R^\vee)$ and $p$ is an LS path. Then
4563: $T$ can be {\em unfolded} in $X$, i.e. there exists a geodesic
4564: triangle $\t{T}\subset X$ such that $g(\t{T})=T$.
4565: \end{thm}
4566: \proof Here is the idea of the proof: We know that billiard
4567: triangles in $\De$ satisfying the {\em simple chain condition} can be
4568: unfolded to geodesic triangles in $X$, see Corollary
4569: \ref{chain-folded}. Littelmann triangle $T$ is billiard, satisfies
4570: the chain condition, but not necessarily the {\em simple chain condition}.
4571: Our goal is to approximate $T$ by Littelmann polygons $P_\eps$,
4572: $\lim_{\eps\to 0} P_\eps=T$, which satisfy the simple chain condition.
4573: We then unfold each $P_\eps$ to a geodesic quadrilateral
4574: $\t{T}_\eps\subset X$. Since $X$ is locally compact, there is a
4575: convergent sequence $\t{T}_{\eps_j}$ whose limit is a geodesic triangle
4576: $\t{T}$ which folds to $T$. Below is the detailed argument.
4577: 
4578: 
4579: 
4580: 
4581: 
4582: 
4583: \begin{figure}[tbh]
4584: \centerline{\epsfxsize=3.5in \epsfbox{unfolding.eps}} \caption{\sl
4585: Approximation.} \label{unfolding.fig}
4586: \end{figure}
4587: 
4588: 
4589: 
4590: 
4591: \medskip
4592: Consider the geodesic path $\pi_\be=\ol{ob}\in {\mathcal P}^+$.
4593: Since $p$ is an LS path with the $\De$-length $\be$, according to
4594: Property 7 in section \ref{operators}, there exists a composition
4595: $\phi\in {\mathcal F}$ of lowering operators so that
4596: $$
4597: \phi(\pi_\be)= p.
4598: $$
4599: 
4600: 
4601: Let $c\subset \De$ denote an alcove which contains the germ of the
4602: segment $\ol{ob}$ at $b$. Pick a point $u$ in the interior of
4603: $\ol{ob}\cap c$. Then for each $\eps>0$ there exists a
4604: point $u_\eps\in int(c)\cap P(R^\vee)\otimes \Q$ such that
4605: 
4606: 
4607: 1. $|u-u_\eps|<\eps$.
4608: 
4609: 
4610: 2. The segments $\ol{ou_\eps}, \ol{u_\eps b}$ do not pass
4611:    through any point of intersection of two or more walls
4612:    (except for the end-points of these segments).
4613: 
4614: 
4615: Observe that $u_\eps$ is a regular point in $(A,W_{aff})$, i.e. it
4616: does not belong to any wall.
4617: 
4618: 
4619: In other words, the path
4620: $$
4621: \hat{p}_\eps:= \ol{ou_\eps}\cup \ol{u_\eps b} \in {\mathcal P}
4622: $$
4623: is {\em generic}. Parameterize $\hat{p}_\eps$ with the constant speed
4624: so that $\hat{p}_\eps(t_\eps)=u_\eps$. See Figure \ref{unfolding.fig}.
4625: 
4626: 
4627: Then the path $\hat{p}_{\eps}$ belongs to ${\mathcal P}^+$; clearly it
4628: is also a generalized LS path: $\hat{p}_\eps\in {\mathcal L}{\mathcal
4629: S}_2$. Moreover,
4630: $$
4631: \lim_{\eps\to 0} \hat{p}_{\eps}= \pi_\be.
4632: $$
4633: Therefore, according to Property 6 of the root operators (see
4634: section \ref{operators}), the operator $\phi$ is defined on all
4635: $\hat{p}_{\eps}$ for $\eps$ sufficiently small and
4636: $$
4637: \lim_{\eps\to 0} \phi(\hat{p}_{\eps})= \phi(\pi_\be)=p.
4638: $$
4639: Set $p_\eps:= \phi(\hat{p}_\eps)$. Since $\hat{p}_\eps$ was generic, the
4640: path $p_\eps$ is generic as well. By construction, for each
4641: sufficiently small $\eps$,
4642: $$
4643: p_\eps(1)=p(1).
4644: $$
4645: Observe also that the germ of the path $p_{\eps}$ at the point
4646: $p_\eps(t_\eps)$ is isomorphic (via an element of
4647: $W_{aff})$ to the germ of $\hat{p}_{\eps}$ at $u_\eps$ (since $u_\eps$
4648: is regular). Similarly, $p_\eps$ is the composition of the path
4649: $p_\eps|[0, t_\eps]$ with the path that belongs to the
4650: $W_{aff}$-orbit of $\ol{u_\eps b}$.
4651: 
4652: 
4653: For each $\eps$ we form a new polygon $P_\eps$ by replacing
4654: the broken side $r(t)=\al+p(t)$ (in $T'$) with the path $\al+p_\eps(t)$.
4655: Clearly,
4656: $$
4657: \lim_{\eps\to 0} P_\eps=T.
4658: $$
4659: 
4660: 
4661: To simplify the notation we now fix $\eps>0$ and let $q:= p_\eps$.
4662: 
4663: 
4664: \begin{lem}
4665: For all sufficiently small $\eps$, the polygon $P_\eps$ is
4666: contained in $\De$.
4667: \end{lem}
4668: \proof Suppose that $\la$ is a simple root which is negative at some
4669: point of the path $\al+q(t)$.
4670: 
4671: 
4672: Since $\la$ is nonnegative on the limiting path $\al+p$, the minimum
4673: of the function $J_\la(t):= \la(q(t)), t\in [0,1]$, converges to zero as $\eps\to 0$.
4674: However, as a generalized LS path, $q$ belongs to ${\mathcal P}_\Z$
4675: (see Property 5 in section \ref{operators}). Since $\al\in
4676: P(R^\vee)$, it follows that the minimum of  $J_\la(t)$ is an
4677: integer. Hence it has to be equal to zero for all sufficiently small
4678: values of $\eps$. Contradiction. \qed
4679: 
4680: 
4681: 
4682: 
4683: Since each $q$ is a generalized LS path and $P_\eps\subset \De$,
4684: the polygon $P_\eps$ is a Littelmann polygon. Moreover, since $q$
4685: is {\em generic}, the polygon $P_\eps$ satisfies the {\em simple
4686: chain condition}. Thus
4687: 
4688: 
4689: 1. For each $t\in [0,t_{\eps})$ either $q$ is smooth at $t$ or
4690: $$
4691: (q'_-(t), q'_+(t))
4692: $$
4693: is a chain of length 1: At $m=q(t)$ the above tangent vectors are
4694: related by a single reflection in $W_{m}$. The fixed-point set of
4695: this reflection is the unique wall passing through $m$.
4696: 
4697: 
4698: 2. The subpath $q([t_\eps,1])$ in $q$ is a geodesic segment and
4699: $$
4700: \del(\eps):= \angle(\ov{u_\eps o}, \ov{u_\eps b})= \pi- \angle(
4701: q'_-(t(\eps)), q'_+(t(\eps))).
4702: $$
4703: 
4704: 
4705: 
4706: 
4707: 
4708: 
4709: Now we are in position to apply Corollary \ref{chain-folded} and
4710: unfold $P_\eps$ in $X$: For each $\eps$ there exists a geodesic
4711: quadrilateral $\t{T}_\eps$ (with one vertex at $o$) in $X$ such that
4712: $$
4713: g(\t{T}_\eps)= P_\eps.
4714: $$
4715: Let $\t{z}=\t{z}_\eps$ denote the point of $\t{T}_\eps$ which maps to $z=q(t_\eps)$
4716: under the folding map $f$. Since $z$ is a regular point, the point
4717: $\t{z}$ is regular as well and the angle between the sides of $\t{T}_\eps$
4718: at $\t{z}$ is the same as the angle between the sides of $P_\eps$ at
4719: $z$, i.e. equals $\del(\eps)$.
4720: 
4721: 
4722: Since the building $X$ is locally compact, the sequence of
4723: quadrilaterals $\t{T}_\eps$ subconverges to a geodesic quadrilateral
4724: $T\subset X$ which is a geodesic triangle since
4725: $$
4726: \lim_{\eps\to 0}\del(\eps)=\pi.
4727: $$
4728: By continuity of the folding $g: X\to\De$,
4729: $$
4730: g(T)= \lim_{\eps\to 0} P_{\eps} =T. \qed
4731: $$
4732: 
4733: 
4734: 
4735: 
4736: In the above proof we assumed that the polygon $T$ is entirely
4737: contained in $\De$. This assumption can be weakened. Let $f: X\to A$
4738: denote the folding $Fold_{a,A}$ into the apartment $A$, where $a$ is
4739: an alcove containing $o$. Suppose that $T\subset A$ is as above, so
4740: that $\al, \ga\in P(R^\vee)$, $p$ is a billiard path, $r=p+\al$. Define two subsets $J, J'\subset I=[0,1]$:
4741: $$
4742: J:= cl( r^{-1}(int( \De))), \quad J':= cl( r^{-1}(int(
4743: V\setminus \De))).
4744: $$
4745: Clearly, $I=J\cup J'$ and the set $J\cap J'$ is finite.
4746: 
4747: 
4748: We assume that for each $t\in J$ the germ of $p$ at $t$ satisfies
4749: the maximal chain condition, and for each $t\in J'$ the germ of $p$
4750: at $t$ is geodesic.
4751: 
4752: 
4753: 
4754: 
4755: \begin{thm}
4756: \label{addendum} Under the above assumptions the polygon $T$ can be
4757: unfolded to a geodesic triangle in $X$ via the retraction $f$.
4758: \end{thm}
4759: \proof Recall that unfolding of $T$ is a local problem of behavior of
4760: the path $r$ at the break-points (Lemma \ref{locality}), which
4761: in our case all occur inside $\De$.
4762: 
4763: 
4764: We first replace $T$ with the polygon $P=\P(T)$, where
4765: $\P=\P_{\De}$ is the projection of $A$ to the Weyl chamber $\De$.
4766: Then, analogously to the proof of Proposition \ref{chaincondition1},
4767: the new polygon $P$ still satisfies the maximal chain condition.
4768: Therefore, according to the previous theorem, the polygon $P$
4769: unfolds in $X$ via the folding map $g=Fold_{\De}: X\to \De$. However
4770: this means that the unfolding condition (stated in Lemma
4771: \ref{locality}) is satisfied at each break-point of the polygon $T$
4772: (since the germs of $r$ and of $\P(r)$ are the same). Hence
4773: the original polygon $T$ unfolds to a geodesic triangle in $X$ via $f: X\to A$. \qed
4774: 
4775: 
4776: 
4777: 
4778: 
4779: 
4780: 
4781: 
4782: 
4783: 
4784: \medskip
4785: Let $\ul{G}$ be a connected split semisimple algebraic group with
4786: the root system $R$ and the cocharacter
4787: lattice $L$ of the maximal torus $\ul{T}\subset \ul{G}$. We let
4788: $\ul{G}^\vee$ denote its Langlands' dual and set
4789: $$
4790: G^\vee:=\ul{G}^\vee(\C).
4791: $$
4792: We assume that $\al,\be,\ga^*\in L$ are dominant weights of $G^\vee$
4793: such that
4794: $$
4795: (V_\al\otimes V_\be \otimes V_{\ga^*})^{G^\vee}\ne 0,$$
4796: equivalently,
4797: $$
4798: V_{\ga}\subset V_\al\otimes V_\be.
4799: $$
4800: 
4801: 
4802: As a corollary of Theorem \ref{LSfolded} we get a new proof of
4803: 
4804: 
4805: \begin{thm}
4806: [Theorem 9.17 in \cite{KLM3}, also proven in \cite{Haines}] \label{klm3}
4807:   Under the above assumptions,
4808: in the thick Euclidean building $X$ there exists a geodesic triangle
4809: with special vertices and the $\De$-side lengths $\al, \be, \ga^*$. In other words,
4810: $$
4811: n_{\al,\be}(\ga)\ne 0 \Rightarrow m_{\al,\be}(\ga)\ne 0.
4812: $$
4813: \end{thm}
4814: \proof Since
4815: $$
4816: V_{\ga}\subset V_\al\otimes V_\be,
4817: $$
4818: according to Littelmann's Theorem \ref{Lit}, there exists a
4819: Littelmann triangle $T'\subset \De$, as in Theorem \ref{LSfolded}.
4820: Let $p\in {\mathcal P}$ denote the LS path (of the $\De$-length
4821: $\be$) representing the broken side of $T'$; $p=\phi(\pi_\be)$,
4822: where $\phi\in {\mathcal F}$ is a composition of lowering operators.
4823: Thus, by Theorem \ref{LSfolded}, there exists a triangle $T=[o, x,
4824: y]\subset X$ such that $Fold_\De(T)=T'$. Therefore, by the
4825: definition of folding,
4826: $$
4827: d_{ref}(o, x)= d_{ref}(o,x')=\al, \quad d_{ref}(o, y)=
4828: d_{ref}(o,y')=\ga.
4829: $$
4830: Since we assumed that $\al, \be\in L\subset P(R^\vee)$ then $x, y$
4831: are special vertices of $x$. Since folding preserves the
4832: $\De$-length,
4833: $$
4834: d_\De(x,y)= \Length(p)=\be. \qed
4835: $$
4836: 
4837: 
4838: 
4839: 
4840: 
4841: 
4842: \subsection{Characterization of folded triangles}
4843: \label{characterizationsection}
4844: 
4845: 
4846: The goal of this section is to extend the results of the previous
4847: one from the case of Littelmann triangles to general broken
4848: triangles satisfying the chain condition.
4849: 
4850: 
4851: 
4852: 
4853: \begin{thm}
4854: \label{chain-unfolding} Suppose that $p\in {\mathcal P}$ is a Hecke
4855: path, $\al=\ov{o u}\in \De$ is such that the path $q:= \al+p$ is
4856: contained in $\De$. Define the billiard triangle $T':=\ol{o u}\cup q
4857: \cup \ol{ q(1) o}$.
4858:  Then $T'$ can be unfolded in $X$.
4859: \end{thm}
4860: \proof The idea of the proof is that the set of unfoldable billiard
4861: paths is closed, thus it suffices to approximate $p$ by unfoldable
4862: paths. We first prove the theorem in the case when $o$ does not
4863: belong to the image of the path $q$.
4864: 
4865: 
4866: 
4867: 
4868: According to Lemma \ref{locality}, unfolding of a path is a purely
4869: local matter. Therefore the problem reduces to the case when $q$ has
4870: only one break-point, $x=q(t_1)$. If $p$ were an LS path, we would
4871: be done. In general it is not, for instance, because it might fail
4872: the {\em maximal chain condition}. We resolve this difficulty by
4873: {\em passing to a smaller Coxeter complex} and a smaller building.
4874: 
4875: 
4876: Let $R_x$ denote the root subsystem in $R$ which is generated by the
4877: roots corresponding to the walls passing through $x$. This root
4878: system determines a Euclidean Coxeter complex where the stabilizer
4879: of the origin is a finite Coxeter group $W'_{sph}$ which is
4880: conjugate to the group $W_x$ via the translation by the vector
4881: $\ov{ox}$. Let $\De_x$ denote the positive Weyl chamber of $(V,
4882: W'_{sph})$ (the unique chamber which contains $\De$). Let $\xi$,
4883: $\eta$ and $\zeta$ denote the normalizations of the vectors
4884: $-p'_-(t_1), p'_+(t_1), \ov{xo}$.
4885: 
4886: 
4887: Then, since $p$ satisfies the chain condition, there exists an
4888: $(S,W_x,\De_x)$-chain
4889: $$
4890: (\nu_0,...,\nu_m), \quad \nu_0=-\xi, \quad \nu_m=\eta, \quad
4891: \nu_i=\tau_i(\nu_{i-1}), \quad 1\le i\le m.
4892: $$
4893: Our first observation is that although this chain may fail to be a
4894: maximal chain with respect to the unrestricted Coxeter complex
4895: $(S,W_{sph})$, we can assume that it is maximal with respect to the
4896: restricted Coxeter complex $(S,W_x)$.
4897: 
4898: 
4899: Next, the initial and final points of $q$ may not belong to
4900: $P(R^\vee_x)$. Recall however that {\em rational points} are dense
4901: in $S$, see Lemma \ref{density}; therefore, there exists a sequence
4902: of rational points (with respect to $R_x$) $\xi_j\in S$ which
4903: converges to $\xi$. Thus, using the same reflections $\tau_i$ as
4904: before, we obtain a sequence of rational chains $(\nu_i^j),
4905: i=0,...,m$, where $\nu_0^j=-\xi_j, \nu_m^j=\eta_j$. We set
4906: $\zeta_j:= \zeta$.
4907: 
4908: 
4909: Hence for each $j$ there exists a number $c=c_j\in \R_+$ so that the
4910: points
4911: $$
4912: x_j=x+ c\xi_j, \quad y_j:= x+ c\eta_j
4913: $$
4914: belong to $P(R^\vee)$.  We define a sequence of paths
4915: $$
4916: q_j:=\ol{x_j x}\cup \ol{x y_j} \in \t{\mathcal P}.
4917: $$
4918: Our next goal is to choose the sequence $\xi_j$ so that the germ of
4919: each $q_j$ at $x$ is contained in $\De$. If $x$ belongs to the
4920: interior of $\De$ then we do not need any restrictions on the
4921: sequence $\xi_j$. Assume therefore that $x$ belongs to the boundary
4922: of $\De$. Let $F$ denote the smallest face of the Coxeter complex
4923: $(V, W_{sph})$ which contains the point $x$ and let $H$ denote the
4924: intersection of all walls through the origin which  contain $x$. It
4925: is clear that $F$ is a convex homogeneous polyhedral cone contained
4926: in $H$ and $x$ belongs to the interior of $F$ in $H$. If $w\in W_x$
4927: is such that $w(-\xi)= \eta$ then $w$ fixes $H$ (and $F$) pointwise.
4928: 
4929: 
4930: By Lemma \ref{density}, applied to the root system $R_x$, there
4931: exists a sequence of unit rational vectors $\xi_j$ and positive
4932: numbers $\eps_j$ converging to zero so that points $x\pm \eps_j
4933: \xi_j$ belong to $int_H(F)$; therefore the sequence $w(x+ \eps_j
4934: \xi_j)$ is also contained in $int_H(F)$.
4935: 
4936: 
4937: Using this sequence $\xi_j$ we define the paths $q_j$; clearly the
4938: germ of $q_j$ at $x$ is contained in $int_H(F)\subset \De\subset
4939: \De_x$.
4940: 
4941: 
4942: \begin{rem}
4943: Note that, typically, the sequence $(c_j)$ is unbounded and the
4944: paths $q_j$ are not contained in $\De_x$.
4945: \end{rem}
4946: 
4947: 
4948: We let $p_j\in {\mathcal P}$ denote the path $q_j- q_j(0)$. Then
4949: each $p_j$ is an LS path with respect to the root system $R_x$:
4950: Integrality and the maximal chain condition now hold. Set
4951: $$
4952: \la_j:= \hbox{length}_{\De_x}(p_j).
4953: $$
4954: 
4955: 
4956: \begin{rem}
4957: Observe that,
4958: $$
4959: \lim_j \bar\la_j = \bar\la\in \De,
4960: $$
4961: where $\la$ is the $\De_x$-length of $p$.
4962: \end{rem}
4963: 
4964: 
4965: Therefore, according to Theorem \ref{addendum} for each $j$ the path
4966: $q_j$ is unfoldable in a thick Euclidean building $X_x$ modeled on
4967: the Coxeter complex
4968: $$
4969: (A, W'_{aff}), \hbox{~~where~~} W'_{aff}= V\ltimes W_x.
4970: $$
4971: This means that there exists a geodesic path $\tilde{q}_j$ in $X_x$,
4972: whose $\De_x$-length is $\la_j$, and which projects to $q_j$ under
4973: the folding $X_x\to A$.
4974: 
4975: 
4976: Let $z_j\in \t{q}_j$ be the points which correspond to the point $x$
4977: under the folding map
4978: $$
4979: \t{q}_j \to q_j.
4980: $$
4981: Thus the ``broken triangle'' $[\xi_j, \zeta, \eta_j]$ in $S_x$
4982: unfolds in $\Si_{z_j}(X_x)$ into a triangle $[\t\xi_j,
4983: \t\zeta_j,\t\eta_j]$ such that
4984: $$
4985: d_{ref}(\t\xi_j, \t\zeta_j) = d_{ref}(\xi_j, \zeta_j),
4986: $$
4987: $$
4988: d_{ref}(\t\eta_j, \t\zeta_j) = d_{ref}(\eta_j, \zeta_j),
4989: $$
4990: $$
4991: d(\t\eta_j,\t\zeta_j)=\pi.
4992: $$
4993: 
4994: 
4995: Observe that the metric distance from $o$ to $z_j$ is uniformly
4996: bounded. Since $X_x$ is locally compact, the sequence of buildings
4997: $\Si_{z_j}(X_x)$ subconverges to the link of a vertex $u\in
4998: X_x\subset X$.
4999: 
5000: 
5001: \begin{rem}
5002: The spherical buildings $\Si_{z_j}(X_x)$, $\Si_{u}(X_x)$ have to be
5003: modeled on the same spherical Coxeter complex $(S,W_x)$, since the
5004: structure group can only increase in the limit and the structure
5005: group at $z_j$ was already maximal possible, i.e. $W_x$.
5006: \end{rem}
5007: 
5008: 
5009: Accordingly, the triangles $[\t\xi_j, \t\zeta_j,\t\eta_j]$
5010: subconverge to a triangle $[\t\xi, \t\zeta,\t\eta]$ whose refined
5011: side-lengths are
5012: $$
5013: d_{ref}(\xi, \zeta),   d_{ref}(\zeta, \eta),\pi.
5014: $$
5015: This shows that the triangle $[\xi,\zeta,\eta]$ can be unfolded in a
5016: building which is modeled on $(S,W_x)$. We now apply the {\em
5017: locality} lemma \ref{locality} to conclude that the path $q$ can be
5018: unfolded in $X$ to a geodesic path. Thus the broken triangle $T'$
5019: unfolds to a geodesic triangle as well.
5020: 
5021: 
5022: \medskip
5023: If $o$ belongs to the image of $q$ we argue as follows. The path
5024: $q$, as before, has only one break point, which in this case occurs
5025: at the origin:
5026: $$
5027: q= \ol{zo}\cup \ol{oy}.
5028: $$
5029: There exists an element $w\in W_{sph}$ which sends the vector
5030: $\eta=-\xi$ to $\xi$, where $\eta$ is the normalization of the
5031: vector $\ov{oy}$. Then consider the geodesic path
5032: $$
5033: \tilde{q}:= \ol{zo}\cup w(\ol{oy}).
5034: $$
5035: It is clear that $g(\tilde{q}) = \P(\tilde{q}) =q$. \qed
5036: 
5037: 
5038: By combining Theorem \ref{chain-unfolding} and Corollary
5039: \ref{folded-chain} we obtain the following
5040: 
5041: 
5042: \begin{thm}
5043: [Characterization of folded triangles]\label{characterization} A
5044: polygon $P\subset \De$ of the form
5045: $$
5046: \ol{ox} \cup (p+\al) \cup \ol{yo}, \hbox{~where~}\al=\ov{ox},
5047: \ga=\ov{oy}\in \De,
5048: $$
5049: can be unfolded to a geodesic triangle in $X$ if and only if $p$ is
5050: a Hecke path, i.e. a billiard path which satisfies the chain
5051: condition.
5052: \end{thm}
5053: 
5054: 
5055: 
5056: 
5057: 
5058: \section{Proof of the saturation theorem}
5059: \label{saturationsection}
5060: 
5061: 
5062: We first prove Theorem \ref{3->4} formulated in the Introduction. Part 1 of Theorem was proven
5063: in \cite{KLM3}, so we prove Part 2.
5064: Let $X$ be a (thick) Euclidean building of rank $r$ modeled on a
5065: discrete Coxeter complex $(A,W)$; the building $X$ is the Bruhat-Tits building associated with
5066: the group $G=\ul{G}(\K)$. Then the assumption  that $m_{\al,\be}(\ga)\ne 0$ is equivalent to
5067: the assumption that there exists a geodesic triangle $T=[\t{x}, \t{y}, \t{z}]\subset X$,
5068: where $\t{x}, \t{y}, \t{z}$ are special vertices of
5069: $X$ and whose $\De$-side-lengths are $\al, \be, \ga^* \in \De\cap P(R^\vee)$.
5070: 
5071: 
5072: Recall that there exists an apartment $\tilde{A}\subset X$ which
5073: contains the segment $\ol{\t{x} \t{y}}$. We let $\t{W}$ denote the
5074: affine Weyl group operating on $\t{A}$. Our first step is to replace
5075: the geodesic triangle $T$ with a geodesic polygon
5076: $$
5077: \tilde{P} := [\tilde z, \t{x}=\tilde{x}_1,..., \tilde{x}_n,
5078: \tilde{x}_{n+1}=\tilde{y}]
5079: $$
5080: as follows. We now treat the point $\t{x}$ as the origin in the
5081: affine space $\t{A}$. Let $\tilde\De\subset \tilde{A}$ be a Weyl chamber in $(\t{A},\t{W})$,
5082: so that $\t\De$ has its tip at $\tilde{x}$ and  $\ol{\t{x}\t{y}}\subset \t\De$.
5083: 
5084: 
5085: Consider the vectors $\varpi_1,...,\varpi_r\in \t\De$ which are the
5086: fundamental coweights of our root system. Then the vector
5087: $\ov{\t{x}\t{y}}$ is the integer linear combination
5088: $$
5089: \ov{\t{x}\t{y}}= \sum_{i=1}^r n_i \varpi_i, n_i\in \N\cup \{0\}.
5090: $$
5091: Accordingly, we define a path $\t{p}$ in $\t\De$ with the initial
5092: vertex $\t{x}$ and the final vertex $\t{y}$ as the concatenation
5093: $$
5094: \t{p}=\pi_{\ul{\la}}= \pi_{\la_1}...*\pi_{\la_r}= \t{p}_1\cup..\cup
5095: \t{p}_r,
5096: $$
5097: where $\la_i:=n_i \varpi_i$, $\ul{\la}=(\la_1,...,\la_r)$. Observe
5098: that the path  $\t{p}$ satisfies the assumptions of Theorem
5099: \ref{chaincondition}. Moreover, $\t{p}$ is a generalized Hecke
5100: path.
5101: 
5102: 
5103: \medskip
5104: Next, let $A\subset X$ be an apartment containing $\ol{\t{z}\t{x}}$,
5105: $a\subset A$ be an alcove containing $\t{z}$; consider the
5106: retraction $f=Fold_{a,A}: X\to A$. This retraction transforms
5107: $\t{P}$ to a polygon $\hat{P}:=f(\t{P})\subset A$ which has geodesic
5108: sides $f(\ol{\t{z}\t{x}})$, $f(\ol{\t{y}\t{z}})$. Note that the
5109: break-points in
5110: $$
5111: \hat{p}:=f(\t{p})= \hat{p}_1\cup...\cup \hat{p}_r,
5112: $$
5113: are the images of the vertices $\t{x}_i$ of $\t{P}$ which are
5114: break-points $\t{p}$, but in addition we possibly have break-points
5115: within the segments $f(\t{p}_i)$. The latter can occur only at the
5116: values of $t$ for which the geodesic segments of $\t{p}_i$ intersect
5117: transversally the walls of $(\t{A},\t{W})$. Since $\t{x}$ is a special vertex and the edges of
5118: each $\t{p}_i$ are parallel to multiples of $\varpi_i$, it follows that each segment $\t{p}_i$ is
5119: contained in the 1-skeleton of $X$. Thus the break-points of $f(\t{p}_i)$ are
5120: automatically vertices of $\t{A}$. We subdivide the path $\t{p}$ so
5121: that all break-points of $\hat{p}$ are the images of the vertices of
5122: $\t{P}$.
5123: 
5124: 
5125: Let $k=k_R$ be the saturation constant of the root system $R$. Then,
5126: according to Lemma \ref{kexists}, for each vertex $v\in A$, the
5127: point $kv\in A$ is a special vertex of $A$. Therefore, applying a
5128: dilation $h\in Dil(A,W)$ with the conformal factor $k$ to the polygon
5129: $\hat{P}$, we obtain a new polygon $k\cdot \hat{P}=h(\hat{P})$,
5130: whose vertices are all special vertices of $A$. Thus we can identify
5131: the Weyl chamber $\De$ with a chamber in $A$ whose tip $o$ is at the
5132: vertex $h(\tilde{z})$ and which contains the geodesic segment
5133: $h(\ol{\tilde{z}\tilde{x}})$.
5134: 
5135: 
5136: Lastly, let $P=\P(k \hat{P})=g(\t{P})$ denote the projection of the
5137: polygon $k\hat{P}$ to the Weyl chamber $\De$, where
5138: $g=Fold_{z,h,\De}$. We set $x:=g(\t{x}), y:= g(\t{y}), x_i:=
5139: g(\t{x}_i)$, $p:= g(\t{p})$, etc.
5140: 
5141: 
5142: \begin{prop}
5143: $P=\ol{ox} \cup p \cup \ol{yo}$ is a Littelmann polygon such that
5144: $$
5145: \length(p)= k\cdot \length(\t{p}).
5146: $$
5147: \end{prop}
5148: \proof We have to show that the path $p$ satisfies the chain
5149: condition with maximal chains at each vertex.
5150: 
5151: 
5152: The chain condition at each vertex of the path $p$ follows
5153: immediately from Theorem \ref{chaincondition}.
5154: 
5155: 
5156: The maximality condition is immediate for the break-points which
5157: occur at the special vertices of $A$, in particular, for all
5158: break-points which are images of the break-points of $\hat{p}$. The
5159: remaining break-points are the ones which occur at the points $\P
5160: (\hat{x}_i)$, where $\hat{x}_i$ are smooth points of $k\hat{p}$ at
5161: which this path transversally intersects the  walls of $A$ passing
5162: through $o$. However at these points the maximality condition
5163: follows from Proposition \ref{chaincondition1}.
5164: 
5165: 
5166: The second assertion of the proposition was proven in Lemma
5167: \ref{invarianceoflength}.  \qed
5168: 
5169: 
5170: \begin{rem}
5171: \label{miniscule}
5172:  Observe that in the case when $\be$ is the sum of
5173: {\em minuscule fundamental coweights}, the multiplication by $k$ in the above
5174: proof is unnecessary since all the vertices of the polygon $\hat{P}$
5175: (and hence $P=\P(\hat{P})$) are already special. Thus, in this case,
5176: the polygon
5177: $$P=Fold_{\De}(\t{P})$$
5178: is a Littelmann polygon.
5179: \end{rem}
5180: 
5181: 
5182: \medskip
5183: The above proposition shows that the polygon $P$ is a Littelmann polygon in $\De$,
5184: which has two geodesic sides having the $\De$-lengths $k\al, k\ga^*$ and
5185: the concatenation of the remaining sides equal to a generalized LS path of the $\De$-length $k\be$.
5186: Therefore, according to Littelmann's theorem (see Theorem
5187: \ref{Lit}),
5188: $$
5189:  V_{k\ga}\subset V_{k\al}\otimes V_{k\be}.
5190: \qed
5191: $$
5192: This concludes the proof of Theorem \ref{3->4}. \qed
5193: 
5194: 
5195: \begin{cor}
5196: \label{corQ3->Q4} Suppose that $\al,\be,\ga\in L\subset P(R^\vee)$
5197: are dominant weights for the complex semisimple Lie group
5198: $\ul{G}^\vee(\C)$, such that $\al+\be+\ga\in Q(R^\vee)$ and that
5199: there exists $N\in \N$ so that
5200: $$
5201:  V_{N\ga}\subset V_{N\al}\otimes V_{N\be}.
5202: $$
5203: Then for the saturation constant $k=k^2_R$ we have
5204: $$
5205: V_{k\ga}\subset V_{k\al}\otimes V_{k\be} .
5206: $$
5207: \end{cor}
5208: \proof Let $X$ be the Euclidean (Bruhat-Tits) building associated to
5209: the group $\ul{G}(\K)$. Then the assumption that
5210: $$
5211:  V_{N\ga}\subset V_{N\al}\otimes V_{N\be}
5212: $$
5213: implies that $(N\al,N\be,N\ga^*)$ belongs to $D_3(X)$ (see
5214: \cite[Theorem 9.17]{KLM3}, or \cite[Theorem 10.3]{KLM3},  or Theorem
5215: \ref{klm3}). Since $D_3(X)$ is a homogeneous cone and $N> 0$, $(\al,\be,\ga^*)\in
5216: D_3(X)$ as well. Moreover, according to Theorem \ref{klm2}, since
5217: $\al, \be, \ga\in P(R^\vee)$ and $\al+\be+\ga\in Q(R^\vee)$, there
5218: exists a triangle $T\subset X$ with the $\De$-side-lengths
5219: $\al,\be,\ga^*$, whose vertices are also vertices of $X$. Now the
5220: assertion follows from Part 3 of theorem \ref{3->4}. \qed
5221: 
5222: 
5223: Using Remark \ref{miniscule} we also obtain:
5224: 
5225: 
5226: \begin{thm}
5227: \label{mini3->4}
5228: Let $X$ be a building as above. Suppose that $T=[x,y,z]$ is a
5229: geodesic triangle in $X$ with the $\De$-side-lengths
5230: $(\al,\be,\ga^*)$, which are dominant weights of $G^\vee$ and so that
5231: one (equivalently, all) vertices of $T$ are special and at least one
5232: of the weights $\al,\be,\ga$ is the sum of minuscule weights. Then
5233: $$
5234: V_{\ga} \subset V_{\al}\otimes V_{\be}.
5235: $$
5236: \end{thm}
5237: 
5238: 
5239: This theorem was originally proven by Tom Haines in the case when
5240: all the weights $\al, \be, \ga$ are sums of minuscules.
5241: 
5242: 
5243: \begin{conjecture}
5244: (T. Haines) Suppose that $\al, \be, \ga$ are sums of minuscule
5245: weights. Then, in the above theorem, the assumption that one vertex
5246: of $T$ is special can be replaced by $\al+\be+\ga\in Q(R^\vee)$.
5247: \end{conjecture}
5248: 
5249: 
5250: Note that (among irreducible root systems) the root systems $G_2, F_4, E_8$ have no minuscule weights,
5251: $B_n, C_n, E_7$ have exactly one minuscule weight and the root systems $A_n, D_n, E_6$ have more than 1 minuscule weights.
5252: For the root system $A_n$ Haines conjecture follows from the saturation theorem. For $D_n$ and $E_6$ it would
5253: follow from the affirmative answer to Question \ref{simplyconjecture}.
5254: 
5255: 
5256: 
5257: \begin{figure}[tbh]
5258: \centerline{\epsfxsize=5in \epsfbox{hecke.eps}} \caption{\sl A Hecke
5259: path which does not satisfy the integrality condition.}
5260: \label{hecke}
5261: \end{figure}
5262: 
5263: \begin{prop}
5264: Suppose that the root system $R$ has exactly one minuscule coweight
5265: $\la$. Then the above conjecture holds for $R$.
5266: \end{prop}
5267: \proof Let $(A,W)$ denote the Euclidean Coxeter complex
5268: corresponding to the root system $R$ and let $X$ be a thick Euclidean building modeled on $(A,W)$.
5269: Given $(\al,\be,\ga^*)\in D_3(X)$ such that $\al, \be, \ga\in P(R^\vee), \al+\be+\ga\in Q(R^\vee)$
5270: we have to construct a geodesic triangle $T=[o,x,y]\subset X$ with special vertices and
5271: the $\De$-side lengths $(\al,\be,\ga^*)$. Clearly, it suffices to treat the case when the root system $R$ is
5272: irreducible and spans $V$. Therefore $R$ has type $B_n, C_n$, or $E_7$. In particular, the index of connection $i$ of $R$ equals $2$
5273: and $-1\in W_{sph}$. In particular, $\ga=\ga^*$.
5274: Let $\la$ denote the unique minuscule coweight of $R$ and let $\La$ denote the span in $\la$ in $V$.
5275: 
5276: Observe that $\la$ does not belong to the coroot lattice $Q(R^\vee)$ and thus, since $i=2$,
5277: $$
5278: \N\cdot \la\cap Q(R^\vee)= 2\N \cdot \la.
5279: $$
5280: Suppose now that $\al=a\la, \be=b\la, \ga=b\la$, where $a, b, c\in \N\cup \{0\}$ and
5281: $$
5282: (\al,\be,\ga)\in D_3(X), \quad \al+\be+\ga\in Q(R^\vee).
5283: $$
5284: Thus $a+b+c$ is an even number and the triple $(a,b,c)$ satisfies
5285: the ordinary metric triangle inequalities.
5286: 
5287: Let $(A',W')=(\R, 2\Z\ltimes \Z/2)$ denote the rank 1 Coxeter complex;  its
5288: vertex set equals $\Z$. The positive Weyl chamber in $(A',W')$ is $\R_+$ and
5289: we can identify $\De'$-distances with the usual metric distances. Let $X'$ denote a thick
5290: building  which is modeled on $(A',W')$ (i.e. a simplicial tree with edges of unit length and thickness $\ge 3$).
5291: Then the above properties of $a, b, c$ imply that $X'$ contains a triangle $T'=[o', x', y']$ with the metric
5292: side-lengths $a, b, c$. If this triangle is contained in a single apartment $A'\subset X'$, we send $T'$ to
5293: a geodesic triangle $T\subset X$ via the isometry $A'\to \La\to A\to X$. If not, we
5294: obtain a folded (Hecke) triangle $P=f'(T')=[o', x', u', f'(y')]\subset \De'$. Note that the unit tangent directions
5295: $\xi', \eta'\in S_{u'}(A')$ to the segments $\ol{u'x'}, \ol{u' f(y')}$ are antipodal. Now embed
5296: the apartment $A'$ into $A\subset X$ via the isometry $\iota$ that sends $A'$ to $\La$, $o'$ to $o$, $1$ to $\la$ (the latter
5297: is a special vertex). Then the point $u:=\iota(u')$ is also a special vertex in $A$.
5298: We claim that the resulting  broken triangle $[o, x, u, y]\subset A$ is a Hecke triangle in $A$.
5299: Indeed, the directions $\xi, \eta$ at $\Si_u$ which are images of $\xi', \eta'$ under $\iota$ are
5300: antipodal and $\eta\in \De$. Therefore, since $u$ is a special vertex and  $-1\in W_{sph}$, according to Lemma
5301: \ref{constructingachain} $\xi\ge \eta$. Thus $\ol{xu}\cup \ol{uy}$ is a Hecke path. It follows that
5302: the broken triangle $[o, x, u, y]$ is a Hecke triangle and hence it unfolds to a geodesic triangle $T$ in $X$.
5303: The triangle $T$  has special vertices and $\De$-side lengths $(\al,\be,\ga)$.
5304: 
5305: Below is an alternative to the above argument. Let $\Si\subset \De^3$ denote the collection of
5306: triples of dominant weights $\tau, \eta, \mu$ such that
5307: $$
5308: (V_\tau \otimes V_\eta \otimes V_\mu)^{G^\vee}\ne 0,
5309: $$
5310: where $G^\vee$ is assumed to be simply-connected.
5311: This set is an additive semigroup, see for instance  \cite[Appendix]{KLM3}.
5312: It suffices to prove that $(\al,\be,\ga)\in \Si$.
5313: 
5314: Set $t:= \frac{1}{2}(a+b-c)$. (The number $t$ is the metric length of the ``leg'' of the geodesic triangle
5315: $T'\subset  X'$ in the above argument, the leg which contains the vertex $x'$.) Since $a+b+c$ is even, the number $t$
5316: is an integer. Set
5317: $$
5318: a_1:= a-t, b_1:= b-t, c_1:= c, \quad \al_1:= a_1\la, \be_1:= b_1\be, \ga_1:= c_1\ga.
5319: $$
5320: Then $c_1=a_1+b_1$ and the metric triangle inequalities for $a, b, c$ imply that $t\ge 0, a_1\ge 0, b_1\ge 0$.
5321: Thus $\al_1, \be_1, \ga_1$ are still dominant weights of $G^\vee$ and they satisfy
5322: $$
5323: \ga_1= \al_1+\be_1.
5324: $$
5325: Then, since $-1\in W_{sph}$ and $\ga_1=\ga_1^*$, we have: $(\al_1,\be_1,\ga_1)\in \Si$. Moreover,
5326: $(t\la, t\la, 0)$ also clearly belongs to $\Si$ and we have
5327: $$
5328: (\al,\be,\ga)= (\al_1,\be_1,\ga_1) + (t\la, t\la, 0)= (\al,\be,\ga). \qed
5329: $$
5330: 
5331: 
5332: 
5333: 
5334: \begin{ex}
5335: There exists a Hecke path $p\in {\mathcal P}$ such that $p(1)\in
5336: P(R^\vee)$, however for the saturation constant $k=k_R$, the path
5337: $k\cdot p$ is not an LS path.
5338: \end{ex}
5339: \proof Our example is for the root system $A_2$, in which case
5340: $k=1$. We will give an example of a Hecke path $p\in {\mathcal P}$
5341: such that $p(1)\in P(R^\vee)$ but $p$ does not belong to ${\mathcal
5342: P}_\Z$. Since, according to Theorem \ref{locint}, each LS path
5343: belong to ${\mathcal P}_\Z$, it proves that $p$ is not an LS path.
5344: 
5345: 
5346: The Hecke path $p$ in question has $\De$-length $\varpi_1+\varpi_2$,
5347: where $\varpi_1, \varpi_2$ are the fundamental coweights; the
5348: break-point of $p$ occurs at the point $-(\varpi_1+\varpi_2)/2$
5349: where the path $p$ backtracks back to the origin.
5350: Thus, for the simple roots $\al$ and $\be$, the minimum of the
5351: functions $\al(p(t)), \be(p(t))$ equals $-1/2$. See Figure
5352: \ref{hecke}.  \qed
5353: 
5354: 
5355: 
5356: 
5357: 
5358: \newpage
5359: \begin{thebibliography}{BaBE}
5360: \addcontentsline{toc}{section}{Bibliography}
5361: 
5362: 
5363: 
5364: 
5365: \bibitem[Ba]{Ballmann}
5366: W.\ Ballmann, ``Lectures on spaces of nonpositive curvature. With an
5367: appendix by Misha Brin.'' DMV Seminar, vol. 25. Birkhauser Verlag,
5368: Basel, 1995.
5369: 
5370: \bibitem[BK]{BK}
5371: P. Belkale, S. Kumar, {\em Eigencone, saturation and Horn problems for symplectic and odd orthogonal groups},
5372:  Preprint arXiv:0708.0398, 2007.
5373: 
5374: \bibitem[BS]{BS}
5375: A.\ Berenstein and R.\ Sjamaar, {\em Coadjoint orbits, moment
5376: polytopes, and the Hilbert-Mumford criterion},  Journ. Amer. Math.
5377: Soc., vol. {\bf 13} (2000), no. 2, p. 433--466.
5378: 
5379: \bibitem[B]{Borel}
5380: A. Borel, ``Linear Algebraic Groups'', Springer Verlag,
5381: Graduate Texts in Mathematics , Vol. 126.
5382: 2nd printing, 1991.
5383: 
5384: 
5385: \bibitem[Bo]{Bourbaki}
5386: N. Bourbaki, ``Lie groups and Lie algebras", Chap. 4, 5, 6. Springer Verlag, 2002.
5387: 
5388: 
5389: \bibitem[Br]{Brown}
5390: K. Brown,``Buildings'', Springer-Verlag, New York, 1989.
5391: 
5392: 
5393: \bibitem[BT]{BT}
5394: F.\ Bruhat\ and\ J. Tits, {\em Groupes reductifs sur un corps
5395: local,} Publ. Math. IHES, No. 41, (1972), p.
5396: 5--251.
5397: 
5398: \bibitem[D]{Demazure}
5399: M. Demazure, {\em Sch\'emas en groupes r\'eductifs},
5400: Bull. Soc. Math. France,  {\bf 93}  (1965) p. 369--413.
5401: 
5402: \bibitem[DW]{DW}
5403: H.\ Derksen and J.\ Weyman, {\em Semi-invariants of quivers and saturation for Littlewood-Richardson coefficients}
5404: J. Amer. Math. Soc.  {\bf 13}  (2000),  no. 3, p. 467--479.
5405: 
5406: 
5407: \bibitem[G]{Garrett}
5408: P. Garrett, ``Buildings and classical groups,'' Chapman \& Hall, London, 1997.
5409: 
5410: 
5411: \bibitem[GL]{GL}
5412: S. Gaussent and P. Littelmann, {\em LS-Galleries, the path model and
5413: MV-cycles},  Duke Math. J.  127  (2005),  no. 1, p. 35--88.
5414: 
5415: 
5416: \bibitem[Gro]{Gross}
5417: B.\ Gross, {\em On the Satake isomorphism}, In: ``Galois
5418: representations in arithmetic algebraic geometry
5419:               (Durham, 1996)'',
5420: London Math. Soc. Lecture Notes, vol. {\bf 254}, (1998) p. 223--237.
5421: 
5422: 
5423: \bibitem[Ha]{Haines}
5424: T.\ J.\ Haines, {\em Structure constants for Hecke and
5425: representations rings}, IMRN vol. {\bf 39} (2003), p. 2103--2119.
5426: 
5427: 
5428: \bibitem[KKM]{KKM}
5429: M. Kapovich, S. Kumar and J. J. Millson, {\em Saturation and irredundancy for $Spin(8)$},
5430: To appear in  Pure and Applied Mathematics Quarterly.
5431: 
5432: \bibitem[KLM1]{KLM1}
5433: M. Kapovich, B.\ Leeb and J.\ J.\ Millson, {\em Convex functions on
5434: symmetric spaces, side lengths of polygons and the stability
5435: inequalities for weighted configurations at infinity}, Preprint,
5436: June 2004.
5437: 
5438: 
5439: \bibitem[KLM2]{KLM2}
5440: M. Kapovich, B.\ Leeb and J.\ J.\ Millson,  {\em Polygons in buildings and
5441: their side-lengths}, Preprint, 2004.
5442: 
5443: 
5444: \bibitem[KLM3]{KLM3}
5445: M. Kapovich, B.\ Leeb and J.\ J.\ Millson,  {\em Polygons in symmetric
5446: spaces and buildings with applications to
5447:   algebra},  Memoirs of AMS, to appear.
5448: 
5449: 
5450: \bibitem[KL]{KleinerLeeb}
5451: B.\ Kleiner and B.\ Leeb, {\em Rigidity of quasi-isometries for
5452: symmetric spaces and Euclidean buildings}, Publ. Math. IHES, vol.
5453: {\bf 86} (1997), p. 115--197.
5454: 
5455: 
5456: 
5457: 
5458: \bibitem[KT]{KnutsonTao}
5459: A. Knutson\ and\ T. Tao, {\em The honeycomb model of ${\rm
5460: GL}_n(\C)$ tensor products. I. Proof of the saturation conjecture,}
5461: J. Amer. Math. Soc., vol. {\bf 12} (1999), no.~4, p. 1055--1090.
5462: 
5463: 
5464: 
5465: 
5466: \bibitem[L1]{Littelmann1}
5467: P.\ Littelmann, {\em  A Littlewood-Richardson rule for symmetrizable
5468: Kac-Moody  algebras},
5469: Invent. Math. 116 (1994), no. 1-3, p. 329--346.
5470: 
5471: 
5472: \bibitem[L2]{Littelmann2}
5473: P.\ Littelmann, {\em Paths and root operators in representation
5474: theory}, Annals of Math. (2)  142 (1995) no. 3, p. 499--525.
5475: 
5476: 
5477: \bibitem[L3]{Littelmann3}
5478: P.\ Littelmann, {\em Characters of representations and paths in
5479: ${\mathfrak H}^*_{\R}$}, In: ``Representation theory and automorphic
5480: forms'' (Edinburgh, 1996), p. 29--49, Proc. Sympos. Pure Math., 61,
5481: Amer. Math. Soc., Providence, RI, 1997.
5482: 
5483: 
5484: \bibitem[Ron]{Ronan}
5485: M. Ronan,``Lectures on buildings'', Perspectives in Mathematics, 7.
5486: Academic Press, Inc., 1989.
5487: 
5488: 
5489: \bibitem[Rou]{Rousseau}
5490: G. Rousseau, {\em Euclidean buildings}, Lectures at \'Ecole d'\'et\'e de
5491: math\'ematiques ``Nonpositively curved geometries, discrete groups and
5492: rigidities'', Institut Fourier, Grenoble, 2004.
5493: 
5494: \bibitem[Sat]{Satake}
5495: I.\ Satake, {\em Theory of spherical functions on reductive algebraic groups
5496: over $p$-adic fields}, Publ. Math. IHES, vol. {\bf 18} (1963), p. 1--69.
5497: 
5498: \bibitem[Sc]{Schwer}
5499: C. Schwer, {\em Galleries, Littlewood polynomials and
5500: structure constants of the spherical Hecke algebra},
5501: Int. Math. Res. Not.  2006, Art. ID 75395, 31 pp.
5502: 
5503: 
5504: \end{thebibliography}
5505: 
5506: 
5507: 
5508: 
5509: 
5510: 
5511: 
5512: 
5513: \bigskip
5514: \noindent Michael Kapovich: \newline Department of Mathematics,
5515: \newline University of California, \newline Davis, CA 95616, USA
5516: \newline
5517:  kapovich$@$math.ucdavis.edu
5518: 
5519: 
5520: \smallskip
5521: \noindent John J. Millson: \newline Department of Mathematics,
5522: \newline University of Maryland, \newline College Park, MD 20742,
5523: USA
5524: \newline
5525:  jjm$@$math.umd.edu
5526: 
5527: 
5528: 
5529: 
5530: \end{document}\end
5531: