math0112251/rbs.tex
1: %&LaTeX
2: %Beginning of AMS-LaTeX source file
3: %
4: % rbs.tex, version 3, 1/7/05
5: %
6: % History of changes:
7: %
8: % version 1, 12/21/01
9: %%
10: %% Changes after version 1:
11: %%
12: %%% Fixed various typos 
13: %%%
14: %%% Fixed the definition of $D_P(V)$ in subsection
15: %%% \ref{ssectVcentralizer} (the previous definition was not independent of
16: %%% the positive system of roots).  The result that needs the corrected
17: %%% definition is Theorem \ref{ssectRaghunathanVanishing}.
18: %%%
19: %%% Fixed the definition of $\n_P(V)$ in subsection
20: %%% \ref{ssectRealSubmodules} (the previous definition was not independent of
21: %%% the positive system of roots)
22: %%%
23: %%% Revised subsections where $D_P(V)$ and \n_P(V) appear:
24: %%% Subsection \ref{ssectIntroVanishingTheoremLmodules}, proof of Lemma
25: %%% \ref{ssectDPVInequality}, proof of Theorem \ref{ssectRaghunathanVanishing},
26: %%% Subsection \ref{ssectRealSubmodules}, statement and proof of The Basic
27: %%% Lemma \ref{ssectBasicLemma}, Corollary \ref{ssectBasicLemmaCor},
28: %%% Subsection \ref{ssectVcentralizerRationalFactors}, proof of Lemma
29: %%% \ref{ssectEqualRankLemma}, Lemma \ref{ssectEqualRankBasicLemmaCor}
30: %%%
31: % version 2, 9/27/02
32: %%
33: %% Changes after version 2:
34: %%
35: %%% Fixed typos
36: %%%
37: %%% Rearranged and better organized topics in Notation subsection,
38: %%% \ref{ssectNotation}
39: %%%
40: %%% Expanded discussion of the Lie algebra cohomology functor of unipotent
41: %%% radicals in Subsubsection \ref{sssectNilpotentCohomology}
42: %%%
43: %%% Revised and expanded treatment of local cohomology with supports (and
44: %%% the Fary and Mayer-Vietoris spectral sequences) in subsection
45: %%% \ref{ssectLocalCohomologyWithSupports} to use $\jhat_Q$ instead of $k$
46: %%% per referee's suggestions. This necessitated several changes throughout
47: %%% the paper.
48: %%%
49: %%% Revised \subsection{Restriction to boundary strata} and the next
50: %%% subsection slightly
51: %%%
52: %%% Revised definition of spectral sequence \eqref{eqnRawEoneTerm}
53: %%%
54: %%% Simplified proof of Proposition
55: %%% \ref{ssectWeightCohomologyLocalCohomologyWithSupports} per referee's
56: %%% suggestion.
57: %%%
58: %%% Moved definition of weak micro-support earlier and characterized it in
59: %%% Theorem \ref{ssectWeightCohomologyMicroSupport}
60: %%%
61: %%% Small changes in exposition
62: %%%
63: %%% Updated references
64: %%%
65: 
66: \documentclass{amsart} \usepackage{amsmath} \usepackage{amscd}
67: \usepackage{amsxtra}
68: 
69: %%% Font Options
70: 
71: \usepackage{amssymb}
72: \usepackage[mathscr]{eucal}	% \mathscr = Euler Script
73: 				% \mathcal = CM calligraphic
74: \DeclareSymbolFont{pssymbols}     {OMS}{ztmcm}{m}{n}
75: \DeclareSymbolFontAlphabet{\mathpsscr}   {pssymbols}
76: 
77: %%% Theorems
78: 
79: \usepackage{amsthm}
80: \swapnumbers
81: \theoremstyle{plain}
82: \newtheorem{thm}[subsection]{Theorem}
83: \newtheorem{cor}[subsection]{Corollary}
84: \newtheorem{prop}[subsection]{Proposition}
85: \newtheorem{lem}[subsection]{Lemma}
86: \newtheorem{sublem}[subsection]{Sublemma}
87: \newtheorem{BasicLemma}[subsection]{The Basic Lemma}
88: \newtheorem*{thm*}{Theorem}
89: \newtheorem*{cor*}{Corollary}
90: \newtheorem*{prop*}{Proposition}
91: \newtheorem*{lem*}{Lemma}
92: \newtheorem*{RapoportConjecturethm}{Theorem \ref{ssectRapoportConjecture}}
93: \newtheorem*{vanthm}{Theorem \ref{ssectGlobalVanishing}}
94: \newtheorem*{SSWHthm}{Theorem \ref{ssectWeightCohomologyMicroSupport}}
95: \newtheorem*{SSIHcor}{Corollary \ref{ssectIHMicroPurityCorollary}}
96: \newtheorem*{SSFibercor}{Corollary
97: \ref{ssectRestrictMicroSupportToFiberCorollary}}
98: \newtheorem*{GoreskyHarderMacPhersonthm}{Theorem \ref{ssectGoreskyHarderMacPhersonTheorem}}
99: 
100: \theoremstyle{definition}
101: \newtheorem{defn}[subsection]{Definition}
102: 
103: \theoremstyle{remark}
104: \newtheorem{rem}[subsection]{Remark}
105: \newtheorem{rems}[subsection]{Remarks}
106: \newtheorem*{rem*}{Remark}
107: \newtheorem*{rems*}{Remarks}
108: \newtheorem{note}[subsection]{Note}
109: \newtheorem*{note*}{Note}
110: 
111: %%% Numbering options
112: 
113: \renewcommand{\theenumi}{\roman{enumi}}
114: \renewcommand{\labelenumi}{(\theenumi)}
115: \newcommand{\itemref}[1]{\textup{(\ref{#1})}}
116: %\numberwithin{equation}{section}
117: \renewcommand{\theequation}{\thesection.\thesubsection.\arabic{equation}}
118: \numberwithin{equation}{subsection}
119: \setcounter{section}{-1}		% start with introduction as section 0
120: \newcounter{saveenum}           % to save and restore interrupted
121: 				% enumerate environments
122: \newcounter{saveenumref}        % to save enumerate counter for modified
123: 				% versions
124: %%% XyPic Options
125: 
126: \usepackage[arrow,matrix,tips,frame,curve,ps,dvips]{xy}
127: %% If dvips is not available, or for device independence at the expense of
128: %% somewhat rougher lines in figures, replace the above line by
129: %% \usepackage[arrow,matrix,tips,frame,curve]{xy}
130: \SelectTips{cm}{}
131: \UseTips
132: \CompileMatrices
133: \def\objectstyle{\displaystyle}
134: \newdir^{ (}{{}*!/-5pt/@^{(}}%	% A new tip style for right hook arrows.
135: 			        % This means the object given by the
136: 			        % directional ^{(} has its reference point
137: 			        % shifted ! by the vector /-5pt/ .
138: \newdir_{ (}{{}*!/-5pt/@_{(}}%	% A new tip style for right hook arrows.
139: 			        % This means the object given by the
140: 			        % directional _{(} has its reference point
141: 			        % shifted ! by the vector /-5pt/ .
142: 
143: %%% Draft manuscript notes
144: 
145: \newcommand{\NB}[1]{ {\bfseries N.B.:} {\itshape #1\/}}
146: 
147: %%% Mathematical Macros
148: 
149: %% coefficient systems
150: 
151: \renewcommand{\AA}{{\mathbb A}}
152: \newcommand{\CC}{{\mathbb C}}
153: \newcommand{\RR}{{\mathbb R}}
154: \newcommand{\ZZ}{{\mathbb Z}}
155: \newcommand{\NN}{{\mathbb N}}
156: \newcommand{\QQ}{{\mathbb Q}}
157: \newcommand{\EE}{{\mathbb E}}
158: \newcommand{\EEtilde}{{\widetilde{\EE}}}
159: \newcommand{\FF}{{\mathbb F}}
160: \newcommand{\HH}{{\mathbb H}}
161: \newcommand{\VV}{{\mathbb V}}
162: 
163: %% operators
164: 
165: \DeclareMathOperator{\Ad}{Ad}
166: \DeclareMathOperator{\codim}{codim}
167: \newcommand{\CoeffDifferential}{\delta}
168: \newcommand{\CoDifferential}{\delta}
169: \newcommand{\ModTwo}{\delta}
170: %	\delta is also used as a root under the macro name \d
171: \DeclareMathOperator{\dom}{Dom}
172: \newcommand{\Dom}{\dom}
173: \DeclareMathOperator{\rank}{rank}
174: \DeclareMathOperator{\Int}{int}
175: \DeclareMathOperator{\Ker}{Ker}
176: \renewcommand{\ker}{\Ker}
177: \renewcommand{\Im}{\operatorname{Im}}
178: \newcommand{\PP}{\mathbb P}
179: \DeclareMathOperator{\Range}{Range}
180: \DeclareMathOperator{\range}{\Range}
181: \DeclareMathOperator{\QQrank}{\QQ-rank}
182: \DeclareMathOperator{\CCrank}{\CC-rank}
183: \DeclareMathOperator{\Span}{Span}
184: \newcommand{\dbar}{\overline{d}}
185: \DeclareMathOperator{\supp}{supp}
186: \newcommand{\geo}{\mathbin{\mathbf o}}
187: \DeclareMathOperator{\Vol}{Vol}
188: \DeclareMathOperator{\mS}{SS}
189: \DeclareMathOperator{\emS}{\mS_{ess}}
190: \newcommand{\essc}{c_{\textup{ess}}}
191: \newcommand{\essd}{d_{\textup{ess}}}
192: \DeclareMathOperator{\Res}{Res}
193: \DeclareMathOperator{\sgn}{sgn}
194: \DeclareMathOperator{\gr}{gr}
195: \DeclareMathOperator{\pr}{pr}
196: \DeclareMathOperator{\Cl}{cl}
197: \newcommand{\cl}[1]{\Cl(#1)}
198: \newcommand{\dd}{d}		% combinatorial part of differential
199: \renewcommand{\l}{\ell}
200: \newcommand{\intprod}{\mathbin{\hbox{\vrule height .5pt width 3.5pt depth 0pt %
201: 	\vrule height 6pt width .5pt depth 0pt}}}
202: \DeclareMathOperator{\restr}{\rho}	% restriction of roots (see also \hsr)
203: \newcommand{\lvvv}{\lVert}				% left boundary norm
204: \newcommand{\rvvv}[1][]{\rVert_{\if!#1!b\else#1,b\fi}}	% right boundary norm
205: 
206: %% relations
207: 
208: \newcommand{\lsb}[1]{{}_{#1}}
209: \newcommand{\lsp}[1]{{}^{#1}\!}
210: \newcommand{\lsbsp}[2]{{}_{#1}^{#2}\!}
211: \newcommand{\tildearrow}{\xrightarrow{\sim}}
212: \newcommand{\longtildearrow}{\mathrel{\overset{\textstyle\mathchar"0366}%
213: {\smash\longrightarrow}}}
214: \newbox\arrowbox
215: \setbox\arrowbox=\hbox{\lower .5ex\rlap{$\longrightarrow$}%
216:  \raise .5ex\hbox{$\longleftarrow$}}
217: \newcommand{\longleftrightarrows}{\mathrel{\text{\copy\arrowbox}}}
218: \newcommand{\mapsfrom}{\leftarrow\mathrel{\mkern-2mu}\mapstochar\,}
219: 
220: %% groups
221: 
222: \newcommand{\G}{\Gamma}
223: \DeclareMathOperator{\SL}{SL}
224: \DeclareMathOperator{\GL}{GL}
225: \DeclareMathOperator{\SO}{SO}
226: \DeclareMathOperator{\Sympl}{Sp}
227: \DeclareMathOperator{\Gal}{Gal}
228: \DeclareMathOperator{\Herm}{Herm}
229: \DeclareMathOperator{\Hom}{Hom}
230: \newcommand{\Pl}{\mathscr P}	% $\Gamma$-conjugacy classes of parabolic
231: 				% $\QQ$-subgroups 
232: 
233: %% spaces
234: 
235: \newcommand{\Dbar}{\overline{D}}
236: \newcommand{\Xbar}{\overline{X}}
237: \newcommand{\Xhat}{\widehat{X}}
238: \newcommand{\back}{\backslash}
239: \newcommand{\Dstar}{D^*}
240: \newcommand{\Xstar}{X^*}
241: \newcommand{\Abar}{{\bar{A}}} % warning: using \overline does not look good
242: 			      % with superscripts to \Abar.
243: \newcommand{\Ybar}{\overline{Y}}
244: \newcommand{\Ftilde}{\widetilde{F}}
245: \newcommand{\X}{\mathcal X}
246: 
247: %% maps
248: 
249: \renewcommand{\i}{i}		% inclusion of closed subset
250: \newcommand{\ihat}{{\hat{\imath}}}
251: \renewcommand{\j}{j}		% inclusion of open subset
252: \newcommand{\jbar}{{\bar{\jmath}}}
253: \newcommand{\jhat}{{\hat{\jmath}}}
254: \DeclareMathOperator{\id}{id}
255: \newcommand{\nil}{p}	% nilmanifold fibration
256: 
257: %% functions
258: 
259: \newcommand{\etahat}{{\hat{\eta}}}
260: 
261: %% neighborhoods
262: 
263: \newcommand{\Wbar}{{\overline{W}}}
264: \newcommand{\What}{{\widehat{W}}}
265: \newcommand{\Uhat}{{\widehat{U}}}
266: 
267: %% sheaves
268: 
269: \newcommand{\C}{{\mathcal C}}
270: \newcommand{\I}{\mathcal I}
271: \newcommand{\IpC}{{\mathcal I_p\mathcal C}}
272: \renewcommand{\H}{\mathcal H}
273: \newcommand{\Dual}{\mathcal D}
274: \newcommand{\Sheaf}{\mathcal S}
275: \newcommand{\A}{\mathcal A}
276: \newcommand{\Asp}{\A_{\,\textup{sp}}}
277: \newcommand{\Acomb}{\A_{\,\textup{comb}}}
278: \renewcommand{\sp}{_{\,\textup{sp}}}
279: \newcommand{\inv}{_{\textup{inv}}}
280: \newcommand{\WnC}{\mathcal W^\eta \mathcal C}
281: \newcommand{\W}{{\mathcal W}}
282: \newcommand{\Gr}{\mathop{\mathcal G r}}
283: \newcommand{\Ecal}{\mathcal E}
284: \newcommand{\F}{\mathcal F}
285: \renewcommand{\L}{\mathscr L}
286: \newcommand{\M}{\mathcal M}
287: 
288: %% cohomology
289: 
290: \newcommand{\IH}{I_{p_w}H}
291: 
292: %% Lie algebras
293: 
294: \newcommand{\sa}{{\mathfrak a}}
295: \newcommand{\h}{{\mathfrak h}}
296: \newcommand{\hb}{{\mathfrak b}}
297: \renewcommand{\k}{{\mathfrak k}}
298: \newcommand{\n}{{\mathfrak n}}
299: \newcommand{\m}{{\mathfrak m}}
300: \newcommand{\levi}{{\mathfrak l}}
301: \newcommand{\p}{{\mathfrak p}}
302: \renewcommand{\r}{{\mathfrak r}}
303: 
304: %% roots and weights
305: 
306: \newcommand{\al}{\alpha}
307: \newcommand{\atilde}{{\widetilde{\alpha}}}
308: \renewcommand{\b}{\beta}
309: \newcommand{\hsr}{\rho}% one-half sum of positive roots (see also
310: 		       % \restr)
311: \renewcommand{\d}{\delta}
312: \newcommand{\dtilde}{{\widetilde{\d}}}
313: \newcommand{\D}{\Delta}
314: \newcommand{\Dhat}{{\widehat{\D}}}
315: \newcommand{\DCC}{\lsb\CC\D}
316: \newcommand{\DRR}{\lsb\RR\D}
317: \newcommand{\e}{\varepsilon}
318: \newcommand{\g}{\gamma}	% We also use \gamma for elements of \Gamma
319: \renewcommand{\u}{\mu}	% \u is used for the highest weight of V.
320: 			% We also use \mu for nilmanifold fibration before
321: 			% the arithmetic quotient.
322: \renewcommand{\v}{\nu}
323: \newcommand{\utilde}{{\widetilde{\u}}}
324: \newcommand{\vtilde}{{\widetilde{\v}}}
325: \newcommand{\ubar}{{\overline{\u}}}
326: \newcommand{\lbar}{{\overline{\lambda}}}
327: \newcommand{\z}{\zeta}
328: \newcommand{\zbar}{{\overline{\zeta}}}
329: \renewcommand{\t}{\tau}
330: \newcommand{\U}{\varUpsilon}
331: \newcommand{\Utilde}{\widetilde{\U}}
332: \newcommand{\Th}{\varTheta}
333: \newcommand{\Phitilde}{\widetilde{\Phi}}
334: \newcommand{\kap}{\kappa}
335: \renewcommand{\o}{\omega}
336: \newcommand{\bd}{\partial}
337: \makeatletter
338: \def\sphat{^{\mathchoice{}{}%
339:  {\,\,\smash[b]{\hbox{\lower4\ex@\hbox{$\m@th\widehat{\null}$}}}}%
340:  {\,\smash[b]{\hbox{\lower3\ex@\hbox{$\m@th\hat{\null}$}}}}}\,}
341: %\newdimen\exa
342: %\exa.2326ex
343: %\renewcommand{\sphat}{^{\mathchoice{}{}%
344: % {\,\,\smash[b]{\hbox{\lower4\exa\hbox{$\mathsurround0pt\widehat{\null}$}}}}%
345: % {\,\smash[b]{\hbox{\lower3\exa\hbox{$\mathsurround0pt\hat{\null}$}}}}}\,}
346: \makeatother
347: 
348: % categories
349: 
350: \newcommand{\Cat}{\mathscr C}	% category
351: \newcommand{\Ob}{\operatorname{Ob}}
352: \newcommand{\Mor}{\operatorname{Mor}}
353: \newcommand{\Rep}{\operatorname{\mathfrak M\mathfrak o\mathfrak d}}
354: \newcommand{\R}{\Rep}
355: \newcommand{\IrrRep}{\operatorname{\mathfrak I\mathfrak r\mathfrak r}}
356: \newcommand{\Sh}{\operatorname{\mathfrak S\mathfrak h}}
357: \newcommand{\Derived}{\operatorname{\mathbf D}}
358: \let\Der\Derived
359: \newcommand{\Complex}{\operatorname{\mathbf C}}
360: \newcommand{\K}{\operatorname{\mathbf K}}
361: \newcommand{\Graded}{\operatorname{\mathbf G\mathbf r}}
362: 
363: %% Miscellaneous
364: 
365: \makeatletter
366: \def\prime{{\null\prime@\null}}
367: \mathchardef\prime@="0230
368: \makeatother
369: 
370: \begin{document}
371: 
372: %%% Top Matter
373: 
374: \author{Leslie Saper}
375: \address{Department of Mathematics\\ Duke University\\ Box 90320\\ Durham,
376: NC 27708\\U.S.A.}
377: \email{saper@math.duke.edu}
378: \urladdr{http://www.math.duke.edu/faculty/saper}
379: \title{$\mathscr L$-modules and micro-support}
380: \thanks{
381: This research was supported in part by the National Science Foundation
382: under grants DMS-8957216, DMS-9100383, and DMS-9870162, a grant from The
383: Duke Endowment, an Alfred P. Sloan Research Fellowship, a grant from the
384: Maximilian-Bickhoff-Stiftung to visit the Katholischen Universit\"at
385: Eichst\"att as Hermann-Minkowski Gastprofessur, and a grant from the
386: Institut des Hautes \'Etudes Scientifiques.  The author
387: wishes to thank these organizations for their hospitality and support.}
388: \thanks{The original manuscript was prepared with the \AmS-\LaTeX\ macro
389: system and the \Xy-pic\ package.}
390: \subjclass{Primary 11F75, 22E40, 32S60, 55N33; Secondary 14G35, 22E45}
391: \keywords{Intersection cohomology, Shimura varieties, locally symmetric
392: varieties, compactifications}
393: \begin{abstract}
394: $\mathscr L$-modules are a combinatorial analogue of constructible sheaves
395: on the reductive Borel-Serre compactification of a locally symmetric space.
396: We define the \emph{micro-support} of an $\mathscr L$-module; it is a set
397: of irreducible modules for the Levi quotients of the parabolic
398: $\QQ$-subgroups associated to the strata.  We prove a vanishing theorem for
399: the global cohomology of an $\mathscr L$-module in term of the
400: micro-support.  We calculate the micro-support of the middle weight profile
401: weighted cohomology and the middle perversity intersection cohomology
402: $\mathscr L$-modules. (For intersection cohomology we must assume the
403: $\QQ$-root system has no component of type $D_n$, $E_n$, or $F_4$.)
404: Finally we prove a functoriality theorem concerning the behavior of
405: micro-support upon restriction of an $\mathscr L$-module to the pre-image
406: of a Satake stratum.  As an application we settle a conjecture made
407: independently by Rapoport and by Goresky and MacPherson, namely, that the
408: intersection cohomology (for either middle perversity) of the reductive
409: Borel-Serre compactification of a Hermitian locally symmetric space is
410: isomorphic to the intersection cohomology of the Baily-Borel-Satake
411: compactification.  We also obtain a new proof of the main result of
412: Goresky, Harder, and MacPherson on weighted cohomology as well as
413: generalizations of both of these results to general Satake
414: compactifications with equal-rank real boundary components.
415: \end{abstract}
416: \maketitle
417: 
418: %%% Table of Contents
419: 
420: \setcounter{tocdepth}{1}
421: \tableofcontents
422: 
423: %%% Introduction
424: 
425: \section{Introduction}
426: \label{sectIntro}
427: Let $X=\Gamma\back D$ be an arithmetic quotient of the symmetric space $D$
428: associated to a connected reductive algebraic group $G$ defined over $\QQ$.
429: (Here $D= G(\RR)/K A_G$, where $K\subseteq G(\RR)$ is a maximal compact
430: subgroup, $A_G=S_G(\RR)^0$, and $S_G$ is the maximal $\QQ$-split torus in
431: the center of $G$.)  If $X$ is not compact, Borel and Serre
432: \cite{refnBorelSerre} construct a compact real analytic manifold with
433: corners $\Xbar$ whose interior is $X$.  The faces of $\Xbar$ are indexed by
434: $\Pl$, the $\G$-conjugacy classes of parabolic $\QQ$-subgroups $P$ of $G$.
435: Each such face $Y_P$ is fibered over an arithmetic quotient
436: $X_P=\G_{L_P}\back D_P$ of the symmetric space $D_P$ associated to the Levi
437: quotient $L_P$ of $P$.  The fibers are compact nilmanifolds (an arithmetic
438: quotient of the real points of the unipotent radical of $P$) and Zucker
439: \cite{refnZuckerWarped} collapses these fibers (for all $Y_P$) to obtain a
440: new compactification $\Xhat$ of $X$, the {\itshape reductive Borel-Serre
441: compactification\/}.
442: 
443: The reductive Borel-Serre compactification is important because it is
444: canonical, its singularities are reasonably easy to understand, and the
445: more singular Satake compactifications $\Xstar$ (such as the
446: Baily-Borel-Satake compactification in the case that $D$ is Hermitian
447: symmetric) may be realized as quotients of it
448: \cite{refnZuckerSatakeCompactifications}.  It would thus be desirable to
449: transfer cohomological calculations from $\Xstar$ to $\Xhat$.  In this
450: paper we develop a combinatorial tool, the theory of $\L$-modules and their
451: micro-support, to study the cohomology of constructible complexes of
452: sheaves on $\Xhat$.  As one application, we settle a 15 year old conjecture
453: on the intersection cohomology made independently by Rapoport
454: \cite{refnRapoportLetterBorel} and by Goresky and MacPherson
455: \cite{refnGoreskyMacPhersonWeighted}.  In fact we prove a generalization:
456: 
457: \begin{RapoportConjecturethm}[Rapoport/Goresky-MacPherson Conjecture]
458: Let $\Xstar$ be a Satake compactification of $X$ for which all real
459: boundary components $D_{R,h}$ are equal-rank, and let $\pi\colon \Xhat \to
460: \Xstar$ be the projection from the reductive Borel-Serre compactification.
461: Let $E$ be a regular representation of $G$ and let $\EE$ be the
462: corresponding locally constant sheaf on $X$.  Let $p$ be a middle
463: perversity.  Then there is a natural quasi-isomorphism of intersection
464: cohomology sheaves $\pi_*\IpC(\Xhat;\EE) \cong \IpC(\Xstar;\EE)$.
465: \end{RapoportConjecturethm}
466: 
467: Recall that $D$ is said to be {\itshape equal-rank\/} if $\CCrank \lsp0 G =
468: \rank K$; here $\lsp 0G$ is the intersection of the kernels of the squares
469: of all characters of $G$ defined over $\QQ$.  The original conjecture was
470: for $D$ Hermitian symmetric and $\Xstar$ the Baily-Borel-Satake
471: compactification.  Previously Zucker had observed that the conjecture held
472: for $G=\Sympl(4)$ with $\EE=\CC$.  Furthermore, Goresky and MacPherson
473: \cite{refnGoreskyMacPhersonWeighted} announced a proof for $G=\Sympl(4)$,
474: $G=\Sympl(6)$, and (for $\EE=\CC$) $G=\Sympl(8)$.  A proof in the case
475: where $\QQrank G=1$ was given by Saper and Stern
476: \cite[Appendix]{refnRapoport}.
477: 
478: The theory of $\L$-modules applies to all $D$ and to sheaves other than
479: intersection cohomology, so for future applications we work in a more
480: general context and do not assume $D$ to be Hermitian or equal-rank except
481: as needed.
482: 
483: The paper is divided into four parts: I. $\L$-modules and Micro-support;
484: II. A Global Vanishing Theorem for $\L$-modules; III. Micro-support
485: Calculations; and IV. Satake Compactifications and Functoriality of
486: Micro-support.  After Part I, Parts II-IV may be read in any order provided
487: one is willing to follow up on a few references; the last section of the
488: paper, however, uses everything that precedes it.  We now briefly outline
489: the main results of the paper, after which we summarize the notation we
490: will use; see also the semi-expository article \cite{refnSaperIHP}.
491: 
492: \subsection{$\L$-modules on the Reductive Borel-Serre
493:   Compactification}
494: (\S\S\ref{sectGeodesicActionBundles}--\ref{sectWeightCohomologyLmodule})\ \
495: The construction of the reductive Borel-Serre compactification $\Xhat=
496: \coprod_P X_P$ is briefly indicated in
497: \S\S\ref{sectGeodesicActionBundles}--\ref{sectReductiveBorelSerre}.  Let
498: $\i_P$ and $\ihat_P$ denote respectively the inclusion of a stratum $X_P$
499: and its closure $\Xhat_P$ into $\Xhat$.  Let $\Derived^b_\X(\Xhat)$ be the
500: derived category whose objects are bounded complexes of sheaves $\Sheaf$ on
501: $\Xhat$ whose cohomology sheaves $\H(\i_P^!\Sheaf)$ supported on each
502: stratum $X_P$ are locally constant, say induced from graded representations
503: $\mathscr E_P$ of $\G_{L_P}$.  We would like to consider such objects
504: equipped with the extra structure that $\mathscr E_P$ is induced by
505: restriction from a graded representation $E_P$ of $L_P$.
506: 
507: In fact it is simpler to start with the family of graded representations
508: $E_P$ of $L_P$ and construct from it an object of $\Derived^b_\X(\Xhat)$.
509: However we need something to ``glue'' these representations together.
510: Consider two strata $X_P$ and $X_Q$ such that $X_P\subseteq \Xhat_Q$, the
511: closure of $X_Q$; this means $P\subseteq Q$ (after perhaps replacing $Q$ by
512: a $\G$-conjugate).  Let $N_P\supseteq N_Q$ be the corresponding unipotent
513: radicals.  The topological link of the stratum $X_P$ when intersected with
514: $X_Q$ is homotopically equivalent to an arithmetic quotient of $N_P^Q(\RR)
515: = N_P(\RR)/N_Q(\RR)$.  A theorem of Nomizu and van Est states that the
516: cohomology of this nilmanifold (with coefficients in $\EE_Q$, the locally
517: constant sheaf associated to $E_Q$) is isomorphic to the Lie algebra
518: cohomology $H(\n_P^Q; E_Q)$; furthermore this is an isomorphism of
519: $\G_{L_P}$-modules.
520: 
521: Thus in \S\ref{sectLsheaves} we are led to consider combinatorial data
522: $\M=(E_\cdot,f_{\cdot\cdot})$ consisting of a family of graded
523: representations $E_P$ of $L_P$, together with ``dual attaching'' morphisms
524: $f_{PQ}\colon H(\n_P^Q;E_Q)\xrightarrow{[1]} E_P$ for all $P\subseteq Q$.
525: If $\M$ is required to satisfy a certain differential type condition
526: \eqref{eqnLsheafCondition} it is called an {\itshape $\L$-module\/}.  From
527: an $\L$-module $\M$ one can construct in a natural way an object
528: $\Sheaf_{\Xhat}(\M)$ of $\Derived^b_\X(\Xhat)$ such that
529: $\H(\i_P^!\Sheaf_{\Xhat}(\M))= \EE_P$.
530: 
531: In fact we may define a category of $\L$-modules $\R(\L_W)$ for any locally
532: closed union of strata $W$ of $\Xhat$.  If $k\colon W\hookrightarrow Z$ is an
533: inclusion of such spaces, we have variants for $\L$-modules of the usual
534: functors $k_*$, $k_!$, $k^*$, and $k^!$, as well as truncation functors by
535: degree and by weight.  In \S\ref{sectRealizationLModules} we see there
536: exists a realization functor $\Sheaf_W\colon \R(\L_W) \to \Derived^b_\X(W)$ and
537: via $\Sheaf_W$ the functors $k_*$, etc. on $\L$-modules correspond to the
538: usual derived functors on $\Derived^b_\X(W)$.  The point is that these
539: derived functors preserve the extra structure.
540: 
541: There are a number of different incarnations of $\Sheaf_W(\M)$ as a complex
542: of sheaves.  In the one we find simplest, a global section of
543: $\Sheaf_W(\M)$ corresponds to a family $\o=(\o_P)$, where $\o_P$ is a
544: special differential form \cite{refnGoreskyHarderMacPherson} on
545: $X_P\subseteq W$ with coefficients in $\EE_P$.  The differential on
546: $\Sheaf_W(\M)$ consists of the exterior derivative on each $\o_P$ together
547: with interaction terms based on the $f_{PQ}$.
548: 
549: If $\M$ is an $\L$-module on $\Xhat$, we define its cohomology
550: $H(\Xhat;\M)$ to be the hypercohomology of the associated complex of
551: sheaves $\Sheaf_{\Xhat}(\M)$.  The complexes of sheaves underlying many
552: cohomology theories on $\Xhat$ can be lifted to $\L$-modules.  For example,
553: let $E$ be a regular representation of $G$.  Then there exist $\L$-modules
554: $\i_{G*}E$, $\IpC(E)$, and $\WnC(E)$ which correspond respectively to the
555: ordinary cohomology $H(X;\EE)=H(\Gamma;E)$, the intersection cohomology
556: $I_pH(\Xhat;\EE)$, and the weighted cohomology $W^\eta H(\Xhat;\EE)$ (see
557: \cite{refnGoreskyHarderMacPherson}).  The latter two $\L$-modules are
558: constructed in \S\S\ref{sectIntersectionCohomologyLmodule} and
559: \ref{sectWeightCohomologyLmodule}.
560: 
561: \subsection{Micro-support}
562: (\S\S\ref{sectMicroSupport}--\ref{sectAlternateMicroSupport})\ \ 
563: To an
564: $\L$-module $\M$ we associate its {\itshape micro-support\/} $\mS(\M)$.
565: The micro-support is a subset of the set of all irreducible representations
566: of $L_P$ for all $P\in \Pl$.  As we will see below, the micro-support of
567: $\M$ limits the range of degrees in which $H(\Xhat;\M)$ can be nonzero.
568: (Micro-support here is vaguely analogous to the micro-support of a sheaf
569: introduced by Kashiwara and Schapira \cite{refnKashiwaraSchapira}, hence
570: the name.  There is no formal correspondence however.)
571: 
572: In order to define $\mS(\M)$, let $V$ be an irreducible $L_P$-module and
573: let $\xi_V$ denote the character by which the maximal $\QQ$-split torus
574: $S_P$ in the center of $L_P$ acts on $V$.  Also let $V|_{M_P}$ denote the
575: restriction of $V$ to $M_P=\lsp0 L_P$, the natural complement to $S_P$.
576: The closed strata $\Xhat_Q$ containing $X_P$ correspond to subsets $\D_P^Q$
577: of $\D_P$, the simple roots of $S_P$ acting on the Lie algebra of the
578: unipotent radical of $P$.  Let $\Xhat_Q\supseteq X_P$ be a closed stratum
579: such that
580: \begin{equation*}
581: \{\,\al\mid (\xi_V+\hsr,\al)< 0\,\} \subseteq \D_P^Q\subseteq  \{\,\al\mid
582: (\xi_V+\hsr,\al)\le  0\,\}
583: \end{equation*}
584: where as usual $\hsr$ denotes one half the sum of the positive roots.  Then
585: $V$ belongs to $\mS(\M)$ if and only if
586: \begin{gather}
587: (V|_{M_P})^* \cong \overline{V|_{M_P}}, \text{ and}
588: \label{eqnMicroSupportConjugateSelfContragradience} \\
589: H(\i_P^*\ihat_Q^!\M)_V\neq 0 \label{eqnMicroSupportVanishing}
590: \end{gather}
591: for some such $Q$.
592: 
593: \subsection{Vanishing Theorem for $\L$-modules}
594: \label{ssectIntroVanishingTheoremLmodules}
595: (\S\S\ref{sectVanishingTheoremLModule}--\ref{sectProofGlobalVanishing})\ \
596: For $V$ in $\mS(\M)$, denote the least and greatest degrees for which
597: \eqref{eqnMicroSupportVanishing} can be nonzero by $c(V;\M)\le d(V;\M)$.
598: Set
599: \begin{align*}
600: c(\M) &= \inf_{V\in\mS(\M)} \tfrac12(\dim D_P - \dim D_P(V)) + c(V;\M)\ ,
601: \text{ and} \\
602: d(\M) &= \sup_{V\in\mS(\M)} \tfrac12(\dim D_P + \dim D_P(V)) + d(V;\M)\ ,
603: \end{align*}
604: where $D_P(V)$ is the symmetric space associated to the reductive
605: $\RR$-subgroup of $L_P$ whose roots (with respect to a fundamental Cartan
606: subalgebra) are orthogonal to the highest weight of $V$.  (This space
607: depends on a $\theta$-stable ordering; we choose one for which $\dim D_P(V)$ is
608: maximized.)  Then we have the
609: 
610: \begin{vanthm}
611: Let $\M$ be an $\L$-module on $\Xhat$. Then  $H^i(\Xhat;\M)=0$ for $i
612: \notin [c(\M), d(\M)]$.
613: \end{vanthm}
614: 
615: The proof of the vanishing theorem proceeds by a combinatorial
616: generalization of the analytic arguments from \cite{refnSaperSternTwo}
617: together with a spectral sequence argument; this reduces the problem to
618: Raghunathan's vanishing theorem for ordinary $L^2$-cohomology, Theorem
619: ~\ref{ssectRaghunathanVanishing}.  In \S\ref{sectLtwo} we recall ordinary
620: $L^2$-cohomology and a version in which restriction to boundary faces is
621: well defined.  Now we need to represent $H(\Xhat;\M)$ by a variant of
622: $L^2$-cohomology for our forms $(\o_P)\in \Sheaf_{\Xhat}(\M)$.  However
623: even though each $\Xhat_P$ is compact this is not entirely straightforward:
624: the $L^2$-norm on $\o_P$ with respect to the natural locally symmetric
625: metric on $\Xhat_P$ is not appropriate because, being a complete metric on
626: a noncompact space, it would impose $L^2$-growth conditions on $\o_P$ which
627: we do not want.  The solution in \S\ref{sectLTwoCohomologyLModules} is to
628: replace $\Xbar$ by a naturally diffeomorphic compact domain $\Xbar_t$
629: within $X$ (this was constructed in \cite{refnSaperTilings}), and use the
630: induced metric; we then work on the corresponding $\Xhat_t$ instead of
631: $\Xhat$.
632: 
633: \subsection{Calculations of
634: Micro-support}(\S\S\ref{sectMicroSupportWeightedCohomology}--\ref{sectPurityProofPartOne})\
635: \ For simplicity, in this introduction we only give the results for the
636: {\itshape essential micro-support\/} $\emS(\M)$; this is the subset of
637: $\mS(\M)$ of $V$ for which \eqref{eqnMicroSupportVanishing} is nonvanishing
638: for all possible $Q$ (in a compatible fashion).  The essential
639: micro-support determines $\mS(\M)$ and may be used in calculating $c(\M)$
640: and $d(\M)$.
641: 
642: For weighted cohomology one can explicitly calculate $H(\i_P^*\ihat_Q^!\M)$
643: and one finds the
644: 
645: \begin{SSWHthm}
646: Let $E$ be an irreducible regular $G$-module and let $\eta$ be a middle
647: weight profile.  The micro-support is nonempty if and only if $(E|_{\lsp0
648: G})^*\cong \overline{E|_{\lsp0 G}}$\textup; if this condition is satisfied
649: then $\emS(\WnC(E))=\{E\}$ and $c(E;\WnC(E))=d(E;\WnC(E))=0$.
650: \end{SSWHthm}
651: 
652: The situation for intersection cohomology is far more delicate.  We do not
653: have an explicit closed formula for $H(\i_P^*\ihat_Q^!\M)$ (or even for the
654: local cohomology $H(\i_P^*\M)$) and we are forced to use an inductive
655: argument.  However condition
656: \eqref{eqnMicroSupportConjugateSelfContragradience} is usually not
657: preserved under passage to a larger parabolic $\QQ$-subgroup.  Nonetheless
658: we have the following result:
659: 
660: \begin{SSIHcor}
661: Assume the irreducible components of the $\QQ$-root system of $G$ are of
662: type $A_n$, $B_n$, $C_n$, $BC_n$, or $G_2$.  Let $E$ be an irreducible
663: regular $G$-module and let $p$ be a middle perversity.  If $(E|_{\lsp0
664: G})^*\cong \overline{E|_{\lsp0 G}}$, then $\emS(\IpC(E))=\{E\}$ with
665: $c(E;\IpC(E))= d(E;\IpC(E))= 0$.
666: \end{SSIHcor}
667: \begin{rems*}
668: Unlike the situation for weighted cohomology, the micro-support can be
669: nonempty even if $(E|_{\lsp0 G})^*\not\cong \overline{E|_{\lsp0 G}}$; see
670: Theorem ~\ref{ssectIHMicroPurity} for details.
671: 
672: The hypothesis on the $\QQ$-root system ensures that the the corresponding
673: $\QQ$-split group for each almost $\QQ$-simple factor of $G$ has a
674: quasi-minuscule representation whose weights are linearly ordered (see the
675: proof of Lemma ~\ref{ssectQInductionLemma}).  We conjecture the results on
676: $\mS(\IpC(E))$ continue to hold without this hypothesis.  The hypothesis is
677: automatically satisfied if $D$ is Hermitian or has a Satake
678: compactification with all real boundary components equal-rank.
679: \end{rems*}
680: 
681: Finally the micro-support of $\i_{G*}E$ is easy to calculate and is treated
682: in \cite{refnSaperIHP}.  As an application of the vanishing theorem it is
683: proved in \cite{refnSaperIHP} that that if $D$ is Hermitian or equal-rank
684: and $E$ has regular highest weight, then $H^i(X;\EE)=0$ for $i<\dim_\CC X$.
685: This answers a question posed by Tilouine.
686: 
687: \subsection{Satake Compactifications}
688: (\S\ref{sectSatakeCompactifications})\ \ We recall the theory of Satake
689: compactifications.  Briefly, given a faithful irreducible finite
690: dimensional representation $\sigma\colon G\to \GL(V)$, Satake
691: \cite{refnSatakeCompact}, \cite{refnSatakeQuotientCompact} constructed a
692: compactification $\lsb\RR \Dstar_\sigma= \coprod_R D_{R,h}$ of $D$, where
693: the strata $D_{R,h}$ are called the {\itshape real boundary components\/}
694: and are indexed by their normalizers $R$.  The union of just the so-called
695: {\itshape rational boundary components\/} yields a space $\Dstar_\sigma$
696: which, under certain conditions on $\sigma$, yields a compactification
697: $\Xstar_\sigma = \G\back \Dstar_\sigma$ of $X=\G\back D$.  The strata $F_R
698: = \G_{L_{R,h}}\back D_{R,h}$ are indexed by a subset
699: $\Pl^\star_\sigma\subseteq \Pl$.  In the case that $D$ is Hermitian
700: symmetric, $\Xstar_\sigma$ for a certain $\sigma$ is the Baily-Borel-Satake
701: compactification.  It has a projective algebraic structure
702: \cite{refnBailyBorel} and its strata are indexed by $\G$-conjugacy classes
703: of maximal parabolic $\QQ$-subgroups.
704: 
705: Zucker showed \cite{refnZuckerSatakeCompactifications} that there is a
706: natural quotient map $\pi\colon \Xhat\to \Xstar_\sigma$.  For each stratum
707: $X_P$ of $\Xhat$ there exists a stratum $F_{P^\dag}$ of $\Xstar_\sigma$
708: such that $\pi|_{X_P}\colon X_P \to F_{P^\dag}$ is a flat bundle with
709: typical fiber $X_{P,\l}$.  In fact, there is a connected normal
710: $\QQ$-subgroup $L_{P,\l}\subseteq L_P$ so that $X_{P,\l}$ is an arithmetic
711: quotient of the symmetric space associated with $D_{P,\l}$, while
712: $F_{P^\dag}$ is an arithmetic quotient of $L_{P,h}= L_P/L_{P,\l}$.  The
713: inverse image $\pi^{-1}(F_R)$ of a stratum $F_R$ of $\Xstar_\sigma$ is the
714: union of those $X_P$ for which $P^\dag=R$; we denote it $X_R(L_{R,\l})$
715: since it is a partial compactification of $X_R$ in the vertical
716: (``$L_{R,\l}$'') directions.  The restriction $\pi|_{X_R(L_{R,\l})}\colon
717: X_R(L_{R,\l})\to F_R$ is again a flat bundle with typical fiber
718: $\Xhat_{R,\l}$.  Let $\ihat_{R,\l}\colon \Xhat_{R,\l}\hookrightarrow
719: X_R(L_{R,\l})$ denote the inclusion of a typical fiber.  In
720: \S\ref{ssectPullbackToFiber} we define functors $\ihat_{R,\l}^*$ and
721: $\ihat_{R,\l}^!$ on $\L$-modules; these commute with their analogues on the
722: derived category.
723: 
724: \subsection{Functoriality of Micro-support}
725: (\S\S\ref{sectEqualRankMicropurityNEW}--\ref{sectRestrictMicroSupportToFiber})\
726: \ We need to understand how micro-support is affected when we apply
727: functorial operations such as $k^*$ or $k^!$ to $\M$.  For brevity we only
728: deal with $k^*$ in this introduction.  A few general results are presented
729: in \S\ref{sectEqualRankMicropurityNEW}.  For the case where $k$ is a closed
730: embedding, however, we only have a result for the {\itshape weak
731: micro-support\/} $\mS_w(\M)$: this is defined similarly to $\mS(\M)$ but
732: omits condition \eqref{eqnMicroSupportConjugateSelfContragradience}.
733: Briefly if an irreducible $L_P$-module $V\in \mS_w(k^*\M)$, then there
734: exists an irreducible $L_{\tilde P}$-module $\tilde V\in \mS_w(\M)$ such
735: that $V$ is an irreducible constituent of $H^\l(\n_P^{\tilde P};\tilde V)$
736: with
737: \begin{equation}
738: (\xi_{V}+\hsr)|_{\sa_P^{\tilde P}} \in \lsp+\sa_P^{\tilde P*}\ ,
739: \label{eqnPositiveWeylHypotheses}
740: \end{equation}
741: where $\lsp+\sa_P^{\tilde P*}$ is the real convex cone generated by
742: $\D_P^{\tilde P}$.
743: 
744: The problem of $\mS_w(k^*\M)$ versus $\mS(k^*\M)$ is eliminated if $k$ is
745: the embedding of $X_R(L_{R,\l})$ for a Satake compactification in which all
746: rational boundary components are equal-rank; see
747: \S\ref{sectFunctorialityMicroSupportEqualRank}.  The point is that if $L_P
748: = \widetilde{L_{P,h}} L_{P,\l}$ is an almost direct product with the
749: symmetric space of $\widetilde{L_{P,h}}$ being equal-rank, then
750: $(V|_{\widetilde{L_{P,h}} })^* \cong \overline{V|_{\widetilde{L_{P,h}}}}$
751: is automatic.
752: 
753: We find that $c(k^*\M)$ and $d(k^*\M)$ may be estimated in terms of $c(\M)$
754: and $d(\M)$, together with the degrees $\l$ from above.  These are the
755: lengths $\l(w)$ of certain Weyl group elements $w$ and
756: \eqref{eqnPositiveWeylHypotheses} gives information on the geometry of the
757: corresponding Weyl chamber.  A basic lemma in \S\ref{sectBasicLemma}
758: translates this geometry into an estimate on $\l(w)$.  In the case that
759: $\#\D_P^{\tilde P} = 1$, the estimate is $\l(w)\le \frac12\dim\n_P^{\tilde
760: P}-\n_P(V)$.  See \S\ref{ssectRealSubmodules} for the definition of
761: $\n_P(V)$; in \S\ref{sectEqualRankBasicLemma} we estimate $\n_P(V)$ in more
762: geometric terms---this requires the stronger condition that all {\itshape
763: real\/} boundary components are equal-rank.
764: 
765: The end result, after restricting to a typical fiber $\Xhat_{R,\l}\subseteq
766: X_R(L_{R,\l})$, is
767: 
768: \begin{SSFibercor}
769: Let $\Xstar_\sigma$ be a Satake compactification of $X$ and assume that all
770: the real boundary components $D_{R,h}$ of the associated Satake
771: compactification $\lsb\RR\Dstar_\sigma$ are equal-rank.  Let $\M$ be an
772: $\L$-module with $\emS(\M)=\{E\}$ for some irreducible regular $G$-module
773: $E$ and $c(E;\M)=d(E;\M)=0$.  Then for every stratum $F_R$ of
774: $\Xstar_\sigma$,
775: \begin{equation*}
776: d(\ihat_{R,\l}^*\M) \le \frac12\codim F_R - \dim \sa_{R}^{G}
777: \qquad\text{and}\qquad c(\ihat_{R,\l}^!\M) \ge \frac12\codim F_R + \dim
778: \sa_{R}^{G}.
779: \end{equation*}
780: \end{SSFibercor}
781: 
782: \subsection{The Conjecture of Rapoport and Goresky-MacPherson} (\S\ref{sectRapoportConjecture})\ \ 
783: Let $\M$ be an $\L$-module.  We wish to identify $\pi_*\Sheaf_{\Xhat}(\M)$
784: as an element of the derived category on $\Xstar_\sigma$.  If $x\in
785: F_R\subseteq \Xstar_\sigma$, then $\pi^{-1}(x) = \Xhat_{R,\l}$, a fiber of
786: $\pi|_{X_R(L_{R,\l})}$.  Thus the local cohomology at $x$ of
787: $\pi_*\Sheaf_{\Xhat}(\M)$ is simply $H(\ihat_{R,\l}^*\M)$, and the local
788: cohomology supported at $x$ is $H(\ihat_{R,\l}^!\M)$.  For $\M = \IpC(E)$,
789: we consequently see from Theorem ~\ref{ssectGlobalVanishing}, Corollary
790: ~\ref{ssectIHMicroPurityCorollary}, and Corollary
791: ~\ref{ssectRestrictMicroSupportToFiberCorollary} that
792: $\pi_*\Sheaf_{\Xhat}(\M)$ satisfies the local characterization of
793: intersection cohomology.  This proves the conjecture of Rapoport and
794: Goresky-MacPherson.  By replacing Corollary
795: ~\ref{ssectIHMicroPurityCorollary} with Theorem
796: ~\ref{ssectWeightCohomologyMicroSupport} we obtain the following
797: generalization of the main theorem of \cite{refnGoreskyHarderMacPherson}:
798: 
799: \begin{GoreskyHarderMacPhersonthm}
800: Under the assumptions of Theorem ~\textup{\ref{ssectRapoportConjecture}},
801: let $\eta$ be a middle weight profile and let $p$ be a middle perversity.
802: Then there is a natural quasi-isomorphism $\pi_*\WnC(\Xhat;\EE) \cong
803: \IpC(\Xstar;\EE)$.
804: \end{GoreskyHarderMacPhersonthm}
805: 
806: \subsection{Work to be Done}
807: In \cite{refnSaperLtwoLmoduleI} the theory of $\L$-modules is generalized
808: to allow $E_P$ to be a locally regular $L_P$-module and hence admit the
809: possibility of non-Hausdorff local cohomology; this allowed us to realize
810: $L^2$-cohomology as an $\L$-module and calculate its micro-support.  There
811: are many other developments and applications of the theory of $\L$-modules
812: that would be interesting to pursue.  We mention for example the following
813: topics: $\QQ$-structures on an $\L$-module and its cohomology (see
814: \S\ref{ssectOtherRealizationsLsheaves}); $\L$-modules on Satake
815: compactifications (for which we use the family
816: $(L_{R,h})_{R\in\Pl^*_\sigma}$ and an appropriate link cohomology
817: functor---see \S\ref{ssectLModuleGeneralizations}); the removal of the
818: hypothesis on the $\QQ$-root system in Theorem ~\ref{ssectIHMicroPurity};
819: the action of Hecke correspondences on $\L$-modules and their cohomology;
820: the ``homotopy category'' of $\L$-modules (see
821: \S\ref{ssectHomotopyLModules}) and its localization with respect to the
822: null system of $\L$-modules with empty micro-support; a ``micro-support''
823: spectral sequence for the cohomology of an $\L$-module; further
824: investigation of the structure of $H(\Gamma;E)$ using $\L$-modules and the
825: relation of this approach to that using Eisenstein series as initiated by
826: Harder (see in particular the work of Franke \cite{refnFranke}).
827: 
828: \subsection{Acknowledgments}
829: I would like to thank Mark Goresky for introducing me to this problem and
830: for discussing it with me and Mark McConnell many times during the academic
831: year 1988-89 while I visited Harvard University.  I would also like to
832: thank Michael Rapoport for reintroducing me to the problem in a letter to
833: Stern and myself in summer 1991, and encouraging us to publish our proof of
834: the $\QQ$-rank 1 case.  I am grateful for stimulating discussions with many
835: people which aided me in this research, in particular, A. Borel, R. Bryant,
836: R. Hain, M. Goresky, G. Harder, J.-P. Labesse, G. Lawler, M. Rapoport,
837: J. Rohlfs, J. Schwermer, J. Tilouine, and N. Wallach.  Finally I would like
838: to thank the referee for many valuable suggestions and corrections, and in
839: particular for suggesting the simplified proof of Proposition
840: ~\ref{ssectWeightCohomologyLocalCohomologyWithSupports} that appears here.
841: 
842: \subsection{Notation}
843: \label{ssectNotation}
844: We now summarize the notation that we will use throughout the paper.
845: 
846: \subsubsection{}
847: The cardinality of a finite set $T$ will be denoted $\#T$.  For $z$ in a
848: group $Z$ and $S\subseteq Z$ a subset, let $\lsp z S = S ^{z^{-1}}$ denote
849: the conjugate $ z S z^{-1}$.  If $Z$ is a topological space, the
850: topological closure of a subset $S\subseteq Z$ will be denoted by $\cl{S}$.
851: The notation $\id_Z$ indicates the identity map of $Z$.
852: 
853: \subsubsection{}
854: \label{sssectNotationCategory}
855: If $\Cat$ is a category we will write $E\in \Cat$ to mean $E$ is an object
856: of $\Cat$.  If $\Cat$ is an additive category, let $\Complex(\Cat)$ denote
857: the category of complexes of objects of $\Cat$ and let $\Graded(\Cat)$
858: denote full subcategory of complexes with zero differential, that is,
859: graded objects of $\Cat$.  The full subcategory of $\Complex(\Cat)$ or
860: $\Graded(\Cat)$ consisting of bounded complexes will be denoted with a
861: superscript $b$.
862: 
863: Let $\Cat$ be an additive category.  For $k\in \ZZ$, the {\itshape shift\/}
864: $C[k]$ of an object $C\in\Complex(\Cat)$ is defined by
865: $C[k]^i=C^{i+k}$ and $d_{C[k]}=(-1)^k d_C$.  Given a functor $F\colon \Cat
866: \to \Complex(\Cat')$, where $\Cat$ and $\Cat'$ are additive categories, we
867: shall often implicitly extend it to a functor $F\colon \Graded(\Cat) \to
868: \Complex(\Cat')$ by setting $F(C)=\oplus_i F(C^i)[-i]$.  Under this
869: convention we have the equality $F(C)[k]=F(C[k])$.  In a few instances we
870: will further extend $F$ to a functor $\Complex(\Cat) \to \Complex(\Cat')$
871: by taking the associated total complex.
872: 
873: \subsubsection{}
874: For a topological space $W$, let $\Sh(W)$ denote the category of sheaves of
875: vector spaces over $\CC$ on $W$.  We write $\Complex(W)$ for
876: $\Complex(\Sh(W))$.  Suppose $W$ is equipped with a stratification $\X$.
877: We will call a complex of sheaves $\Sheaf$ \emph{constructible}%
878: %
879: \footnote{In the terminology of \cite[V,
880:     3.3(ii)]{refnBorelIntersectionCohomology} this is $\X$-cohomologically
881:   locally constant, or $\X$-clc.}
882: %
883: if for all strata $X_P\in \X$, the restriction $H(\Sheaf)|_{X_P}$ of the
884: local cohomology sheaves is locally constant.  Let $\Complex_\X(W)$ denote
885: the category of complexes of sheaves on $W$ which are constructible.  Let
886: $\Derived_\X(W)$ be the associated derived category, obtained by formally
887: inverting quasi-isomorphisms in the corresponding homotopy category.  We
888: denote the usual derived functors on $\Derived_\X(W)$ without the prefix
889: ``$R$\/'', for example, $i_*$ instead of $Ri_*$.
890: 
891: \subsubsection{}
892: All references to manifolds, smooth differential forms, fiber
893: bundles, locally constant sheaves, etc. should be taken in the sense of
894: $V$-manifolds \cite{refnSatakeVManifold} (also called a orbifolds
895: \cite{refnThurston}).
896: 
897: For $Y$ a manifold with (possibly empty) boundary or corners and $\EE$ a
898: locally constant sheaf on $Y$, let $\EE\mapsto \A(Y;\EE)$ denote the de
899: Rham resolution functor consisting of sheaves of smooth differential forms
900: on $Y$ with values in $\EE$.  The differential is exterior differentiation,
901: denoted $d=d_Y=d_{Y,\EE}$.  Note that if $\EE$ is graded we follow the sign
902: convention of \S\ref{sssectNotationCategory}.  The complex of global
903: sections is denoted $A(Y;\EE)$ and the subcomplex of forms with compact
904: support is denoted $A_c(Y;\EE)$.
905: 
906: \subsubsection{}
907: For any algebraic group defined over $\RR$, we denote the Lie algebra of
908: its real points by the corresponding Gothic letter, for example, $\mathfrak
909: g = \operatorname{Lie} G(\RR)$.  The complexification is denoted by a
910: subscript $\CC$, for example $\mathfrak g_{\CC}$.  Now let $Z$ be an
911: algebraic group defined over $\QQ$.  The group of characters of $Z$ defined
912: over $\QQ$ is denoted $X(Z)$.  The one-dimensional representation space
913: associated to a character $\chi$ is denoted $\CC_\chi$.  Let $\lsp 0Z =
914: \bigcap_{\chi\in X(Z)} \ker \chi^2$ be the intersection of the kernels of
915: the squares of characters of $Z$.  The group $Z$ is an {\itshape almost
916: direct product\/} of $\QQ$-subgroups $H$ and $K$ if the product morphism
917: $H\times K\to Z$ is surjective with finite kernel.
918: 
919: \subsubsection{}
920: For a connected reductive $\QQ$-group $L$, let $\R(L)$ denote the category
921: of regular representations of $L$.  We write $\Complex(L)$ and $\Graded(L)$
922: instead of $\Complex(\R(L))$ and $\Graded(\R(L))$.  If $M\subseteq L$ is a
923: reductive $\QQ$-subgroup and $E\in \R(L)$, let $E|_M$ denote the regular
924: representation of $M$ obtained from $E$ by restriction.
925: 
926: Let $\IrrRep(L)$ denote the isomorphism classes of irreducible regular
927: $L$-modules.  For any $E\in \R(L)$ we have an {\itshape isotypical
928: decomposition\/}
929: \begin{equation*}
930: E=\bigoplus_{V\in\IrrRep(L)} E_V
931: \end{equation*}
932: where $E_V=\Hom_{L}(V,E)\otimes V$.  For any irreducible representation
933: $V\in\IrrRep(L)$, we let $\xi_V$ denote the character by which $S_L$ acts
934: on $V$.  Since $S_L$ is $\QQ$-split, $\xi_V$ is defined over $\QQ$.
935: Although $\xi_V$ may not extend to a character of $L$, there exists a
936: natural number $k$ such that $\xi_V^k$ extends to a character of $L$
937: defined over $\QQ$.
938: 
939: \subsubsection{}
940: Let $Z$ be an algebraic group and let $\sigma\colon Z \to \GL(V)$ be a
941: regular representation.  Recall the contragredient representation
942: $\sigma^*$ on $V^*$ is defined by $(\sigma^*(g)\psi)(v)=
943: \psi(\sigma(g^{-1})v)$.  Define the complex vector space $\overline V$ to
944: have vectors $\overline v$ indexed by $v\in V$, with operations $\overline
945: v + \overline w = \overline{v+w}$ and $\lambda \overline v =
946: \overline{\overline \lambda v}$.  If $Z$ is defined over $\RR$ there is a
947: complex conjugate representation $\overline\sigma$ on $\overline V$ defined
948: by $\overline\sigma(g)\overline v = \overline{\sigma(\overline g)v}$.
949: 
950: \subsubsection{}
951: In this paper $G$ will be a connected reductive algebraic group defined
952: over $\QQ$, and $\G\subseteq G(\QQ)$ will be an arithmetic subgroup.  (We
953: briefly depart from this convention on $G$ in
954: \S\S\ref{ssectSatakeCompactificationsSymmetricSpaces},
955: \ref{ssectRealBoundaryComponents}.)  We let $D$ denote the associated
956: symmetric space $G(\RR)/KA_G$, where $K$ is a maximal compact subgroup of
957: $G(\RR)$ and $A_G$ is the connected component of the group of real points
958: of the maximal $\QQ$-split torus $S_G$ in the center of $G$.  Let
959: $X=\G\back D$ be the arithmetic quotient.  If $E$ is a regular
960: representation of $G$, the induced locally constant sheaf on $X$ will be
961: denoted by the corresponding blackboard bold letter, thus $\EE=D\times_{\G}
962: E$.
963: 
964: Since we do not assume $\G$ to be neat, $X$ (and the various other
965: arithmetic quotients that we will consider) need not be manifolds, but
966: rather $V$-manifolds.  Alternatively the reader may assume that $\G$ is neat; any
967: arithmetic group contains a neat subgroup with finite index.
968: 
969: \subsubsection{}
970: \label{sssectSplitComponents}
971: Let $P$ be a parabolic $\QQ$-subgroup of $G$, with unipotent radical $N_P$
972: and Levi quotient $L_P=P/N_P$. Set $M_P= \lsp 0 L_P$.  Let $S_P$ denote the
973: maximal $\QQ$-split torus in the center of $L_P$ and set $A_P =
974: S_P(\RR)^0$.  Restriction of characters yields an injective morphism
975: $X(L_P)\to X(S_P)$ whose image has finite index.  In terms of real points
976: we have $L_P(\RR) = M_P(\RR) \times A_P$.
977: 
978: If $P\subseteq Q$ are parabolic $\QQ$-subgroups, $P/N_Q$ is a parabolic
979: $\QQ$-subgroup of $L_Q$ with unipotent radical $N_P^Q=N_P/N_Q$.  One can
980: naturally identify $S_Q$ as a subgroup of $S_P$ and $X(L_Q)$ as a subgroup
981: of $X(L_P)$.  Namely consider the diagram
982: \begin{equation*}
983: \xymatrix{
984: {L_P} & {P/N_Q} \ar@{>>}[l]_-{j} \ar@{^{ (}->}[r]^-{i} & {L_Q\ ;}
985: }
986: \end{equation*}
987: since $P/N_Q$ is parabolic in $L_Q$, $i$ identifies its center with that of
988: $L_Q$, while $j$ identifies it with a subgroup of the center of $L_P$.
989: Furthermore
990: \begin{equation*}
991: \xymatrix{
992: {X(L_P)} \ar[r]^-{\sim} & {X(P/N_Q)}  & {X(L_Q)\ .}
993: \ar@{_{ (}->}[l]_-{i^*}
994: }
995: \end{equation*}
996: 
997: Set $S_P^Q = \bigl( \bigcap_{\chi\in X(L_Q)} \ker \chi\cap S_P\bigr)^0$.
998: Then $S_P$ is an almost direct product
999: \begin{equation*}
1000: S_P = S_Q \cdot S_P^Q  \ .
1001: \end{equation*}
1002: If we write $A_P^Q= S_P^Q(\RR)^0$, then $A_P$ is an honest direct product
1003: \begin{equation}
1004: A_P = A_Q \times A_P^Q  \ .  \label{eqnUsualDecomp}
1005: \end{equation}
1006: 
1007: \subsubsection{}
1008: \label{sssectSplitComponentRoots}
1009: Let $P$ be a parabolic $\QQ$-subgroup of $G$.  Let $\sa_P$ be the Lie
1010: algebra of $A_P$ and let $\langle\ ,\ \rangle$ denote the natural pairing
1011: $\sa_P^*\times\sa_P\to \RR$.  We may identify $X(S_P)\otimes_\ZZ \RR\cong
1012: \sa_P^*$.  Consequently we will view elements of $X(S_P)$ both as
1013: characters on $S_P$ and as linear functionals on $\sa_P$.
1014: 
1015: The group $L_P$ acts via conjugation on $N_P$ (and hence on $\n_{P\CC}$)
1016: via any lift $\widetilde{L_P}\subseteq P$.  Although this action depends on
1017: the lift, the characters $\al\in X(S_P)$ by which $S_P$ acts are
1018: independent of the lift.  We denote these characters by $\Phi(\n_P,S_P)$ or
1019: $\Phi(\n_P,\sa_P)$.  As usual we call them ``roots'' even though in general
1020: for $P$ nonminimal they are not the positive roots of a root system.  They
1021: do, however, have the property that every such root is a unique nonnegative
1022: integral linear combination of so-called simple roots; we let $\D_P$ denote
1023: these simple roots.
1024: 
1025: In the case that $P=P_0$ is a minimal parabolic $\QQ$-subgroup, $\D_{P_0}$
1026: is a basis of the $\QQ$-root system $\lsb\QQ\Phi$ of $G$.  In this case we
1027: omit the subscript $P_0$ and simply write $\D$, $S$, $\sa$, etc.
1028: 
1029: \subsubsection{}
1030: If $P\subseteq Q$ are parabolic $\QQ$-subgroups, the {\itshape type of $Q$
1031: with respect to $P$\/} is the subset $\D_P^Q$ of $\D_P$ consisting of roots
1032: restricting to 1 on $S_Q$.  Parabolic $\QQ$-subgroups $Q$ containing $P$
1033: are in one-to-one correspondence with subsets of $\D_P$ via
1034: $Q\leftrightarrow \D_P^Q$.
1035: 
1036: Let
1037: \begin{equation*}
1038: \sa_P = \sa_Q \oplus \sa_P^Q
1039: \end{equation*}
1040: be induced from \eqref{eqnUsualDecomp}.  By \S\ref{sssectSplitComponents}
1041: we have a natural inclusion $\sa_Q^* \hookrightarrow \sa_P^*$.  Then
1042: $\sa_Q$ is the subspace of $\sa_P$ annihilated by $\D_P^Q$, whereas
1043: $\sa_P^Q$ is the subspace annihilated by $\D_Q$ together with a basis of
1044: $X(S_G)$.  We also have a dual decomposition
1045: \begin{equation*}
1046: \sa_P^* = \sa_Q^* \oplus \sa_P^{Q*}\ .
1047: \end{equation*}
1048: Given $\al\in \sa_P^*$ we write $\al = \al_Q + \al^Q$ for its corresponding
1049: decomposition and likewise for elements of $\sa_P$.
1050: 
1051: \subsubsection{}
1052: For each $\g\in \D$, let $\g\spcheck\in \sa$ be the corresponding coroot.
1053: (We shall occasionally identify $\sa$ with $\sa^*$ via a Weyl group
1054: invariant inner product $(\ ,\ )$, in which case $\g\spcheck =
1055: \frac{2}{(\g,\g)}\g$.)  To define coroots where $P$ is an arbitrary
1056: parabolic $\QQ$-subgroup, we proceed as in \cite{refnArthurTraceFormula}.
1057: Namely for $\al \in \D_P$, let $\g\in \D\setminus \D^P$ be the unique root
1058: restricting to $\al$ on $\sa_P$ and let $\al\spcheck= (\g\spcheck)_P$ be
1059: the projection of $\g\spcheck$ to $\sa_P$.
1060: 
1061: If $P\subseteq Q$, then $\sa_P^Q$ has a basis of coroots,
1062: $\{\al\spcheck\}_{\al\in\D_P^Q}$.  Let
1063: $\Dhat_P^Q=\{\b_\al^Q\}_{\al\in\D_P^Q}$ be the dual basis of $\sa_P^{Q*}$.
1064: On the other hand, $\D_P^Q$ is a basis for $\sa_P^{Q*}$.  For $\b\in
1065: \Dhat_P^Q$, define $\b\spcheck\in \sa_P^Q$ by $\langle
1066: \al,\b\spcheck\rangle = \langle \b,\al\spcheck\rangle$ for all $\al \in
1067: \D_P^Q$.  Then $\{\b\spcheck\}_{\b\in \Dhat_P^Q}$ is the basis of $\sa_P^Q$
1068: dual to $\D_P^Q$.
1069: 
1070: \subsubsection{}
1071: \label{sssectRootAndDominantCones}
1072: Let $\lsp +\sa_P^*$ denote the real convex cone spanned by $\D_P$ and let
1073: $\sa_P^{*+}$ denote the real convex cone spanned by $\Dhat_P$.  Then
1074: $\sa_P^{*+}$ is the dominant cone
1075: \begin{equation*}
1076: \{\,\u\in\sa_P^*\mid \langle\u,\al\spcheck\rangle\ge 0
1077: \text{ for all $\al\in\D_P$}\,\}
1078: \end{equation*}
1079: and $\sa_P^{*+} \subseteq \lsp +\sa_P^*$.  Similarly define $\lsp
1080: +\sa_P^{Q*}$ and $\sa_P^{Q*+}$.
1081: 
1082: \subsubsection{}
1083: Beginning in \S\ref{sectSatakeCompactifications} we will need to consider
1084: parabolic $\RR$-subgroups.  The discussion in
1085: \S\S\ref{sssectSplitComponents}--\ref{sssectRootAndDominantCones} applies
1086: with $\QQ$ replaced by $\RR$.  We will add a left subscript $\RR$ to the
1087: notation in this case, for example, $\lsb\RR S_P$, $\lsb\RR A_P$,
1088: $\lsb\RR\D_P$, etc.
1089: 
1090: \subsubsection{}
1091: Let $P\subseteq Q$ be parabolic $\QQ$-subgroups.  Let $(P,Q)$ denote the
1092: parabolic $\QQ$-subgroup ``complementary'' to $Q$ with respect to $P$ with
1093: type $\D_P\setminus \D_P^Q$.  When $(P,Q)$ occurs as a subscript or a
1094: superscript, we omit the parentheses, for example, $\D_P^{P,Q}$.  Thus we
1095: have a disjoint decomposition
1096: \begin{equation*}
1097: \D_P = \D_P^{P,Q} \amalg \D_P^Q\ .
1098: \end{equation*}
1099: Restriction to $\sa_Q$ sends $\D_P^Q$ to zero and defines a bijection
1100: $\D_P^{P,Q}\tildearrow \D_Q$.
1101: 
1102: Note that this decomposition of $\D_P$ induces a direct sum decomposition
1103: \begin{equation*}
1104: \sa_P = \sa_Q \oplus \sa_{P,Q}^G\ ,
1105: \end{equation*}
1106: however in general $\sa_{P,Q}^G \neq \sa_P^Q$.
1107: 
1108: \subsubsection{}
1109: Let
1110: \begin{equation*}
1111: \Pl=  \{\, \text{$\G$-conjugacy classes of parabolic $\QQ$-subgroups of
1112: $G$} \,\}\ ;
1113: \end{equation*}
1114: notationally we do not distinguish a parabolic $\QQ$-subgroup from its
1115: $\G$-conjugacy class.  Define a partial order on $\Pl$ by
1116: \begin{equation*}
1117: P \le Q \quad \Longleftrightarrow \quad P \subseteq Q^\gamma \text{ for
1118: some $\gamma\in\G$.}
1119: \end{equation*}
1120: (When $P\le Q$ we will always feel free to replace $Q$ by a $\G$-conjugate
1121: so as to arrange $P\subseteq Q$ as needed.)  For $P$ and $Q$ in $\Pl$, let
1122: $P\vee Q$ denote the least upper bound of $P$ and $Q$; let $P\cap Q$ denote
1123: the greatest lower bound when it exists.
1124: 
1125: Given $P\le R$, let $[P,R]$ denote the interval $ \{\, Q\in \Pl \mid P\le Q
1126: \le R\,\}$.  This is a complemented lattice: for every $Q\in [P,R]$ there
1127: exists $Q'\in [P,R]$ such that $Q\cap Q' = P$ and $Q\vee Q' = R$, namely
1128: $Q' = (P,Q)\cap R$.  We represent this situation by the diagram
1129: \begin{equation}
1130: \vcenter{\xymatrix @ur @M=1pt @R=1.5pc @C=1pc {
1131: {Q'} \ar@{-}[r] \ar@{-}[d] & {R} \ar@{-}[d] \\
1132: {P\rlap{\qquad .}} \ar@{-}[r] & {Q}
1133: }} \label{eqnParallelogram}
1134: \end{equation}
1135: 
1136: \subsubsection{}
1137: Let $P$ be a parabolic $\QQ$-subgroup and let $\h$ be a Cartan subalgebra
1138: of $\levi_P$.  Write $\h=\hb_P \oplus\sa_P$ where $\hb_P$ is a Cartan
1139: subalgebra of $\m_P$.  (In \S\ref{sectEqualRankBasicLemma} we shall also
1140: need $\h=\lsb\RR\hb_P\oplus\lsb\RR\sa_P$ for $P$ a parabolic $\RR$-subgroup,
1141: where $\lsb\RR\hb_P$ is a Cartan subalgebra of $\lsb\RR\m_P$.)  The
1142: corresponding roots will be denoted $\Phi(\levi_{P\CC},\h_\CC)$ and a
1143: positive system will be denoted $\Phi^+(\levi_{P\CC},\h_\CC)$.  (In the
1144: case $Q=G$ we shall often simply write $\Phi$ and $\Phi^+$.)
1145: 
1146: Via a lift, $\h$ acts on $\n_P$ via the adjoint action; the corresponding
1147: roots are independent of the lift and will be denoted
1148: $\Phi(\n_{P\CC},\h_{P\CC})$.  Likewise via a lift, $\h$ may be considered
1149: as a Cartan subalgebra of $\levi_Q$ for any parabolic $\QQ$-subgroup
1150: $Q\supseteq P$; the decomposition $\h=\hb_Q+\sa_Q$ and the roots
1151: $\Phi(\levi_{Q\CC},\h_\CC)$ are independent of the lift.  We can always
1152: choose a positive system $\Phi^+(\levi_{Q\CC},\h_\CC)$ containing
1153: $\Phi^+(\levi_{P\CC},\h_\CC)$ and $\Phi(\n_{P\CC},\h_{P\CC})$.
1154: 
1155: \subsubsection{}
1156: Let $\hsr = \hsr^P + \hsr_P$ be one-half the sum of roots in
1157: $\Phi^+(\mathfrak g_\CC,\h_\CC)$, where $\hsr^P$ (resp. $\hsr_P$) is
1158: one-half the sum of the roots in $\Phi^+(\levi_{P\CC},\h_\CC)$
1159: (resp. $\Phi(\n_{P\CC},\h_\CC)$).  We may identify $2\hsr_P$ with the
1160: $\QQ$-rational character of $L_P$ by which $L_P$ (via any lift $\widetilde
1161: L_P$) acts on $\bigwedge^{\dim \n_P} \n_P$.  Thus $\hsr_P$ is supported on
1162: $\sa_P$ and $\hsr=\hsr^P+\hsr_P$ corresponds to the decomposition induced
1163: by restriction to the summands in $\h=\hb_P \oplus \sa_P$.
1164: 
1165: \subsubsection{}
1166: \label{sssectNilpotentCohomology}
1167: Let $P$ be a parabolic $\QQ$-subgroup and let $E$ be a regular $G$-module.
1168: The \emph{nilpotent Lie algebra cohomology functor} $E\mapsto H(\n_P;E)$
1169: from $\R(G)$ to $\Graded(L_P)$ will play a crucial role in what follows.
1170: As a vector space, $H(\n_P;E)$ is defined as the cohomology of the complex
1171: $C(\n_P;E)=\Hom(\bigwedge\n_P,E)$ equipped with the usual differential (see
1172: for example \cite[I, \S1.1]{refnBorelWallach}).
1173: %
1174: %define $d\colon C(\n_P;E) \to C(\n_P;E)[1]$ by
1175: %\begin{multline*}
1176: %d\o(X_0, \dots,X_p) = \sum_i (-1)^i X_i\cdot \o(X_0,\dots,\hat
1177: %X_i,\dots, X_p) + \\
1178: %\sum_{i<j} (-i)^{i+j} \o([X_i,X_j],X_0,\dots,\hat
1179: %X_i,\dots,\hat X_j, \dots, X_p).
1180: %\end{multline*}
1181: %Then $d^2=0$
1182: %
1183: The coadjoint representation induces an action of $P$ on $C(\n_P;E)$ which
1184: commutes with the differential, and hence there is an action of $P$ on
1185: $H(\n_P;E)$.  The unipotent radical $N_P$ acts trivially \cite[I,
1186: \S1.1(5)]{refnBorelWallach} so this descends to a representation of $L_P$.
1187: Finally, a morphism of $G$-modules $f\colon E \to E'$ induces a morphism of
1188: $P$-modules $C(\n_P;E) \to C(\n_P;E')$ which commutes with the
1189: differentials and hence induces an $L_P$-module morphism $H(\n_P;E) \to
1190: H(\n_P;E')$ which we denote $H(\n_P;f)$.
1191: 
1192: \subsubsection{}
1193: Let $W$ denote the Weyl group of $\Phi$.  For $w\in W$ let $\l(w)$ denote
1194: the length of $w\in W$ with respect to the reflections in simple roots and
1195: set $\Phi_w = \{\,\g\in \Phi^+\mid w^{-1}\g < 0\,\}$.  It is  well known
1196: that $\l(w) = \#\Phi_w$.
1197: 
1198: For a parabolic $\QQ$-subgroup $P$ let $W^P\subseteq W$ be the subgroup
1199: corresponding to the  Weyl
1200: group of $\Phi(\levi_{P\CC},\h_\CC)$ and let $W_P\subseteq W$ be the set of
1201: minimal length coset representatives of $W^P\backslash W$.%
1202: %
1203: \footnote{Our use of subscript versus superscript here is opposite from the
1204: usual convention, but this way is more compatible with the subscripts and
1205: superscripts on $A$ and $\Delta$.}
1206: %
1207: We have $w\in W_P$ if and only if $\Phi_w \subseteq \Phi(\n_{P\CC},\h_\CC)$.
1208: There is a factorization
1209: \begin{equation*}
1210: W=W^PW_P.
1211: \end{equation*}
1212: 
1213: The corresponding groups and coset representatives for the $\QQ$-root
1214: system will be used in \S\ref{sectInductionLemmaProof}, and will be denoted
1215: there $\lsb\QQ W$, etc.
1216: 
1217: \subsubsection{Kostant's Theorem}
1218: \label{ssectKostantsTheorem}
1219: Let $E$ be a regular $G$-module.  The complement $W_P$ plays an important
1220: role in Kostant's theorem \cite{refnKostant}, which states that $H(\n_P;E)$
1221: has a decomposition as an $L_P$-module indexed by $W_P$:
1222: \begin{equation*}
1223: H(\n_P;E) = \bigoplus_{w\in W_P} H(\n_P;E)_w \ .
1224: \end{equation*}
1225: Furthermore, $H(\n_P;E)_w$ is concentrated in degree $\l(w)$ and, if $E$ is
1226: irreducible with highest weight $\lambda$, then $H(\n_P;E)_w$ is the
1227: irreducible $L_P$-module having highest weight $w(\lambda+\hsr)-\hsr$.
1228: 
1229: \subsubsection{}
1230: \label{ssectKostantDegeneration}
1231: Let $P\subseteq Q$ be parabolic $\QQ$-subgroups.  We have
1232: \begin{equation*}
1233: W = W^QW_Q =(W^PW_P^Q)W_Q
1234: \end{equation*}
1235: and one may check that
1236: \begin{equation*}
1237: W_P=W_P^QW_Q.
1238: \end{equation*}
1239: For $w\in W_P$ write $w=w^Qw_Q$ accordingly.  We have the important
1240: equality
1241: \begin{equation*}
1242: \l(w)= \l(w^Q) + \l(w_Q)
1243: \end{equation*}
1244: and in fact
1245: \begin{equation}
1246: \Phi_w=\Phi_{w^Q} \textstyle \coprod w^Q \Phi_{w_Q}   \ .
1247: \end{equation}
1248: It immediately follows from Kostant's theorem that
1249: \begin{equation*}
1250: H(\n_P;E)_w \cong H(\n_P^Q;H(\n_Q;E)_{w_Q})_{w^Q}
1251: \end{equation*}
1252: and consequently \cite[(11.8)]{refnGoreskyHarderMacPherson},
1253: \cite[4.10]{refnSchwermerGeneric}
1254: \begin{equation}
1255: H(\n_P;E) \cong H(\n_P^Q;H(\n_Q;E))\ .
1256: \label{eqnKostantDegeneration}
1257: \end{equation}
1258: 
1259: %%% Body of Paper
1260: 
1261: \specialsection*{Part I. $\L$-modules and Micro-support}
1262: 
1263: \section{Geodesic Action and Related Bundles}
1264: \label{sectGeodesicActionBundles}
1265: Recall that $D$ is the symmetric space $G(\RR)/KA_G$ associated to $G$ and
1266: $X=\G\back D$ is an arithmetic quotient.  In this section we briefly recall
1267: the geodesic retraction and nilmanifold fibration associated to a parabolic
1268: $\QQ$-subgroup $P$ \cite[\S3]{refnBorelSerre},
1269: \cite[\S3]{refnZuckerWarped}, \cite[ \S\S3,~ 4,
1270: ~7]{refnGoreskyHarderMacPherson}.  This section also serves to establish
1271: our notation for the various spaces and bundles that will enter into the
1272: construction of the reductive Borel-Serre compactification in
1273: \S\ref{sectReductiveBorelSerre}.  One novelty of our presentation is that
1274: we introduce a ``geodesic action'' for the entire Levi quotient $L_P(\RR)$
1275: as opposed to merely $A_P$; this is convenient for describing the flat
1276: connection on the nilmanifold fibration.
1277: 
1278: \subsection{Geodesic action}
1279: \label{ssectGeodesicAction}
1280: Given any basepoint $x\in D$, let $L_{P,x}$ denote the unique lift of $L_P$
1281: to $P$ which is stable under the Cartan involution associated to $x$.
1282: Since $P(\RR)$ acts transitively on $D$, we may express $y\in D$ as $px$
1283: for $p\in P(\RR)$; since $P=N_P\rtimes L_{P,x}$ we may write $p = u r$ with
1284: $u\in N_P(\RR)$ and $r\in L_{P,x}(\RR)$.  Define the {\itshape geodesic
1285: action\/} of $L_P(\RR)$ on $D$ by
1286: \begin{equation}
1287: z \geo y = u z_x r x, \label{eqnGeodesicAction}
1288: \end{equation}
1289: where $z_x\in L_{P,x}(\RR)$ denotes the lift of $z$.  Note that at the
1290: basepoint $x$ the geodesic action of an element of $L_P$ agrees with the
1291: ordinary action of its lift in $L_{P,x}$.  Clearly the restriction of this
1292: action to $A_P$ agrees with the geodesic action as defined in
1293: \cite{refnBorelSerre}.
1294: 
1295: \begin{lem*}
1296: The geodesic action $(z,y)\mapsto z\geo y$ of $L_P(\RR)$ on $D$ is
1297: well-defined \textup(that is, $z\geo y$ is independent of the choice of $p$
1298: and $x$ above\textup).  It satisfies
1299: \begin{equation}
1300: q(z\geo y) = \lsp q z\geo qy \qquad (q\in P(\RR)), \label{eqnGeodesicAndUsual}
1301: \end{equation}
1302: where $z\mapsto \lsp q z$ denotes the action of $P(\RR)$ by conjugation on
1303: $L_P(\RR)$.
1304: \end{lem*}
1305: 
1306: \begin{proof}
1307: Consider first the choice of $p$.  If $y =
1308: p'x$ for $p'\in P(\RR)$, then $p'=pk$ for some $k\in K_x\cap P\subseteq
1309: L_{P,x}$ and hence $p'=u r'$ with $r'=r k$.  Thus $u z_x r' x = u
1310: z_x r x$ and so the right hand side of \eqref{eqnGeodesicAction} depends only
1311: on $x$, $y$, and $z$.
1312: 
1313: Now consider another basepoint $x'$.  Write $x'=qx = vl x$ with $v\in
1314: N_P(\RR)$ and $l\in L_{P,x}(\RR)$.  Then $L_{P,x'}=vL_{P,x}v^{-1}$.  We may
1315: express $y= p'x'$ where $p'= pq^{-1}= u r l^{-1} v^{-1}= (uv^{-1})(vr
1316: l^{-1} v^{-1}) = u' r'$.  Thus we compute $u' z_{x'}r'x'
1317: =(uv^{-1})(vz_xv^{-1})(vr l^{-1} v^{-1})qx = u z_x r x$, which shows that
1318: \eqref{eqnGeodesicAction} is independent of $x$.
1319: 
1320: To prove \eqref{eqnGeodesicAndUsual}, calculate $q(z\geo y) = quz_xrx = \lsp q u
1321: \lsp q(z_x) \lsp qr qx$.  However $\lsp q(z_x)$ is the lift of $\lsp qz$ to
1322: $L_{P,x'}$, where $x'=qx$.  Thus $q(z\geo y) =
1323: \lsp q u (\lsp qz)_{x'} \lsp qr x' = \lsp qz \geo (\lsp q u  \lsp qr x') =
1324: \lsp qz \geo qy$.
1325: \end{proof}
1326: 
1327: \begin{cor*}
1328: The geodesic action of $L_P(\RR)$ and the usual action of $N_P(\RR)$
1329: commute.  The geodesic action of $A_P\subseteq L_P(\RR)$ commutes with the
1330: usual action of $P(\RR)$.
1331: \end{cor*}
1332: 
1333: \begin{proof}
1334: By the lemma we need to show that $ \lsp q z=z$ for $q\in N_P(\RR)$ and
1335: $z\in L_P(\RR)$ (and likewise for $q\in P(\RR)$ and $z\in A_P$).  Write
1336: $z=z_x N_P(\RR)$ for some basepoint $x$.  Then $\lsp q z = qz_x
1337: q^{-1}N_P(\RR) =z_x q^{z_x}q^{-1}N_P(\RR)$, so we need to show that
1338: $q^{z_x}q^{-1}\in N_P(\RR)$.  This follows in the first case since $q\in
1339: N_P(\RR)$ and $N_P(\RR)$ is normal.  In the second case, write $q=vl$ with
1340: $v\in N_P(\RR)$ and $l\in L_{P,x}(\RR)$.  Then $q^{z_x}q^{-1} =v^{z_x}
1341: l^{z_x} l^{-1} v^{-1} = v^{z_x}v^{-1}$, since $z_x\in A_{P,x}$ is in the
1342: center of $L_P(\RR)$.  Now proceed as in the first case.
1343: \end{proof}
1344: 
1345: \subsection{Comparison of geodesic actions}
1346: \label{ssectCompareGeodesicActions}
1347: Consider now two parabolic $\QQ$-subgroups $P\subseteq Q$; as in
1348: \S\ref{sssectSplitComponents} the split component $A_Q$ may be viewed as a
1349: subgroup of $A_P$.  Thus it has a geodesic action viewed as a subgroup of
1350: either $L_Q(\RR)$ or $L_P(\RR)$; these two geodesic actions agree
1351: \cite[3.10]{refnBorelSerre}.
1352: 
1353: For the extended geodesic action we have the
1354: 
1355: \begin{lem*}
1356: The geodesic action of $N_P^Q(\RR) = N_P(\RR)/N_Q(\RR)
1357: \subseteq L_Q(\RR)$ commutes with the geodesic action of $L_P(\RR)$.
1358: \end{lem*}
1359: 
1360: \begin{proof}
1361: Write $y\in D$ as $y=urx$, where $u\in N_P(\RR)$, $r\in L_{P,x}(\RR)$.
1362: Since $N_P \subseteq Q = N_Q
1363: \rtimes L_{Q,x}$ we can decompose $N_P= N_Q \rtimes N_{P,x}^Q$ where
1364: $N_{P,x}^Q\subseteq L_{Q,x}$ is the lift of $N_P^Q$.  Thus we can
1365: write $u =vw$, where $v\in N_Q(\RR)$ and $w\in
1366: N_{P,x}^Q(\RR)$.  The geodesic action of $n\in N_P^Q(\RR)$ is now $n\geo y
1367: = v n_x wrx$, which clearly commutes with the geodesic action of $z\in
1368: L_P(\RR)$, $z\geo y = u z_x r x= vwz_x r x$.
1369: \end{proof}
1370: 
1371: \subsection{Geodesic retraction}
1372: \label{ssectGeodesicRetraction}
1373: By Corollary ~\ref{ssectGeodesicAction} the geodesic action of $A_P$ and
1374: the ordinary action of $\lsp0 P(\RR)$ commute.  The consequent action of
1375: $A_P\times \lsp0 P(\RR)$ on $D$ is transitive and the stabilizer of any
1376: point $x$ is $\{A_G\}\times\lsp0 P(\RR)\cap K_x$, where $K_x\subseteq
1377: G(\RR)$ is the maximal compact subgroup fixing $x$.  Set%
1378: %
1379: \footnote{In \cite{refnBorelSerre} $e_P$ is denoted $e(P)$.}
1380: %
1381: $\mathscr A_P^G = \lsp0 P(\RR)\back D$ and $e_P=A_P\back D$.  The preceding
1382: implies there is an isomorphism
1383: \begin{equation}
1384: D \cong \mathscr A_P^G \times e_P \label{eqnTrivializationPrincipalAPBundle}
1385: \end{equation}
1386: of $(A_P^G\times \lsp0 P(\RR))$-homogeneous spaces.  The quotient map
1387: \begin{equation*}
1388: D\longrightarrow e_P=A_P\back D
1389: \end{equation*}
1390: obtained by collapsing the orbits of the geodesic action of $A_P$ is a
1391: trivial principal $A_P^G$-bundle with canonical trivializing sections given
1392: by the orbits of $\lsp0 P(\RR)$.  Note that after choosing a basepoint
1393: $x\in D$ and hence $\bar x\in \mathscr A_P^G$, there is an isomorphism of
1394: $A_P^G$-homogeneous spaces $A_P^G\tildearrow \mathscr A_P^G$ given by
1395: $a\mapsto a\geo \bar x$.  Thus $\mathscr A_P^G$ is an affine version of
1396: $A_P^G$.
1397: 
1398: The action of $\G_P=\G\cap P$ on $D \cong \mathscr A_P^G \times e_P$
1399: operates only on the second factor, so the principal $A_P^G$-bundle
1400: structure persists after taking the quotient:
1401: \begin{equation*}
1402: r_P\colon  \G_P\back D \longrightarrow Y_P=\G_P\back e_P.
1403: \end{equation*}
1404: The map $r_P$ is called {\itshape geodesic retraction}.  Since $\G_P\subset
1405: \lsp0 P(\RR)$, the quotients by $\G_P$ of the orbits of $\lsp0 P(\RR)$
1406: still yield canonical trivializing sections of $r_P$.
1407: 
1408: If $P\subseteq Q$ this map descends to yield a canonically trivial
1409: $A_P^Q$-bundle (also called geodesic retraction)
1410: \begin{equation*}
1411: r_P\colon  \G_P\back e_Q \longrightarrow Y_P.
1412: \end{equation*}
1413: 
1414: \subsection{Nilmanifold fibrations}
1415: \label{ssectNilmanifoldFibrations}
1416: The geodesic action of $\lsp 0L_P(\RR)$ descends to $e_P$ and by Corollary
1417: ~\ref{ssectGeodesicAction} this action commutes with the $N_P(\RR)$-action.
1418: The resulting action of $N_P(\RR)\times \lsp 0L_P(\RR)$ on $e_P$ is
1419: transitive and the stabilizer of any point is contained in $\{1\}\times
1420: \lsp 0L_P(\RR)$.  Set $\mathscr N_P(\RR)= \lsp 0L_P(\RR)\back e_P$ and $D_P
1421: = N_P(\RR)\back e_P$; the space $\mathscr N_P(\RR)$ is an affine version of
1422: $N_P(\RR)$ and $D_P$ is the generalized symmetric space associated to
1423: $L_P$.  Then there is an isomorphism
1424: \begin{equation*}
1425: e_P\cong \mathscr N_P(\RR)\times D_P
1426: \end{equation*}
1427: of $(N_P(\RR)\times \lsp 0L_P(\RR))$-homogeneous spaces. The principal
1428: $N_P(\RR)$-bundle
1429: \begin{equation*}
1430: \mu\colon  e_P \longrightarrow D_P
1431: \end{equation*}
1432: obtained by collapsing the orbits of the action of $N_P(\RR)$ has canonical
1433: trivializing sections given by the orbits of the geodesic action of $\lsp
1434: 0L_P(\RR)$.  Given a basepoint $x\in D$ and hence $\bar x\in\mathscr
1435: N_P(\RR)$, there is an isomorphism $n_x\colon  N_P(\RR)\tildearrow\mathscr
1436: N_P(\RR)$ of $N_P(\RR)$-homogeneous spaces defined by $n_x(u)= u \bar x$.
1437: 
1438: The action of $\lsp 0P(\RR)$ on $e_P$ descends to actions on both $\mathscr
1439: N_P(\RR)$ and $D_P$, and the isomorphism $e_P\cong \mathscr N_P(\RR)\times
1440: D_P$ is an isomorphism of $\lsp 0P(\RR)$-homogeneous spaces.  Set $\mathscr
1441: N_P(\RR)' = \G_{N_P}\back \mathscr N_P(\RR)$ and note that the action of
1442: $\G_P$ on $\mathscr N_P(\RR)$ descends to an action of
1443: $\G_{L_P}=\G_P/\G_{N_P}$ on $\mathscr N_P(\RR)'$.  The previous
1444: decomposition descends to an isomorphism $\G_{N_P}\back e_P \cong \mathscr
1445: N_P(\RR)' \times D_P$ of $\G_{L_P}$-homogeneous spaces and thus there is a
1446: canonically trivial $\mathscr N_P(\RR)'$-bundle
1447: \begin{equation*}
1448: \mu'\colon  \G_{N_P}\back e_P \to D_P\ .
1449: \end{equation*}
1450: Set%
1451: %
1452: \footnote{In \cite{refnGoreskyHarderMacPherson} the group $N_P(\RR)$ is
1453: denoted $\mathpsscr U_P$ and the nilmanifold $N_P(\RR)'$ is denoted $N_P$.
1454: What we denote here as $\mathscr N_P(\RR)$ and $\mathscr N_P(\RR)'$ are not
1455: explicitly mentioned.}
1456: %
1457: $N_P(\RR)' = \G_{N_P}\back N_P(\RR)$.  Given a basepoint $x\in D$, the
1458: isomorphism $n_x$ descends to $n_x'\colon  N_P(\RR)'\tildearrow \mathscr
1459: N_P(\RR)'$.
1460: 
1461: Finally take the quotient by $\G_{L_P}$ of $\mu'$; we obtain a flat bundle
1462: (with fibers noncanonically diffeomorphic to $\mathscr N_P(\RR)'$)
1463: \begin{equation*}
1464: \nil\colon  Y_P \longrightarrow X_P = \G_{L_P}\back D_P.
1465: \end{equation*}
1466: This is simply the flat bundle $ D_P \times_{\G_{L_P}} \mathscr N_P(\RR)'
1467: \to X_P$ associated to the $\G_{L_P}$-space $\mathscr N_P(\RR)'$. It is
1468: called the {\itshape nilmanifold fibration}.
1469: 
1470: \subsection{}
1471: \label{ssectActionOnNilmanifoldGivenBasepoint}
1472: For later use, we calculate \cite[(7.8)]{refnGoreskyHarderMacPherson} the
1473: action of $\G_{L_P}$ on $N_P(\RR)'$ that is needed in order to make $n_x'\colon 
1474: N_P(\RR)'\tildearrow \mathscr N_P(\RR)'$ an isomorphism of
1475: $\G_{L_P}$-spaces.  Let $\g\G_{N_P}\in \G_{L_P}$ and let $\G_{N_P}u\in
1476: N_P(\RR)'$; write $\g=vl$ where $v\in N_P(\RR)$ and $l\in L_{P,x}(\RR)$.
1477: Then
1478: \begin{equation*}
1479: \g\G_{N_P}\cdot n_x'(\G_{N_P}u) = \G_{N_P} \g. u.\bar x = \G_{N_P}
1480: v.l.u.\bar x = \G_{N_P} v.\lsp lu. \bar x = n_x'(\G_{N_P}v.\lsp lu)\ .
1481: \end{equation*}
1482: Thus the induced action of $\g$ on $N_P(\RR)'$ is $\G_{N_P}u \mapsto
1483: \G_{N_P}v.\lsp lu = \G_{N_P}\lsp \g u. v$.
1484: 
1485: \section{The Reductive Borel-Serre Compactification}
1486: \label{sectReductiveBorelSerre}
1487: Let $X=\G\back D$ be a locally symmetric space.  We describe the
1488: Borel-Serre compactification $\Xbar$ \cite{refnBorelSerre} and its quotient
1489: introduced by Zucker \cite{refnZuckerWarped}, the reductive Borel-Serre
1490: compactification $\Xhat$ \cite[\S8]{refnGoreskyHarderMacPherson}.
1491: 
1492: \subsection{The Borel-Serre compactification}
1493: \label{ssectBorelSerre}
1494: Let $P$ be a parabolic $\QQ$-subgroup of $G$.  There is an isomorphism
1495: $A_P^G \tildearrow (\RR^{> 0})^{\D_P}$ defined by $a\mapsto
1496: (a^\al)_{\al\in\D_P}$ and thus $A_P^G$ may be embedded into a semigroup
1497: $\Abar_P^G\tildearrow (\RR^{> 0}\cup\{\infty\})^{\D_P}$.
1498: 
1499: The orbits of the action of $A_P^G$ on $\Abar_P^G$ yield a stratification
1500: indexed by the parabolic $\QQ$-subgroups $Q\supseteq P$.  Namely, let
1501: $o_Q\in\Abar_P^G$ be defined by $\al(o_Q)=\infty$ for $\al\in\D_P^{P,Q}$,
1502: and $\al(o_Q) = 1$ for $\al\in\D_P^Q$.  Then we have the stratification
1503: \begin{equation}
1504: \Abar_P^G = \coprod_{Q\supseteq P} A_P^G\cdot o_Q.
1505: \label{eqnAbarStratification}
1506: \end{equation}
1507: 
1508: Recall that the quotient map $D \to e_P$ by the geodesic action of $A_P^G$
1509: is a principal $A_P^G$-bundle.  Let $D(P)= D \times_{A_P^G} \Abar_P^G$ be
1510: the associated $\Abar_P^G$-bundle.  The orbits of $A_P^G\times \lsp0
1511: P(\RR)$ yield a stratification
1512: \begin{equation*}
1513: D(P) = \coprod_{Q\supseteq P} e_Q\ .
1514: \end{equation*}
1515: If $P\subseteq Q$ there is a natural identification $D(Q)\subseteq D(P)$ as
1516: an open union of strata.
1517: 
1518: Given the above identifications, set $\Dbar = \bigcup_P D(P)$.  This is a
1519: manifold-with-corners whose stratification by faces is
1520: \begin{equation*}
1521: \Dbar =  \coprod_{P} e_P\ .
1522: \end{equation*}
1523: The construction of $\Dbar$ may be generalized to apply to $e_P$ (being a
1524: space of type $S-\QQ$ \cite[2.3, 3.9]{refnBorelSerre}) and the resulting
1525: $\overline{e_P}$ may be identified with the closure $\cl{e_P}$ of $e_P$ in
1526: $\Dbar$.
1527: 
1528: The action of $G(\QQ)$ extends to $\Dbar$ and the quotient $\Xbar=\G\back
1529: \Dbar$ is compact.  This is the {\itshape Borel-Serre compactification\/}
1530: of $X$.  On each stratum $e_P$ the equivalence relation induced by $\G$ is
1531: that of $\G_P=\G\cap P$, and two strata $e_P$ and $e_{P'}$ become
1532: identified in the quotient if and only if $P'=\lsp \g P$ for some $\g\in
1533: \G$.  Thus the strata of $\Xbar$ are indexed by the $\G$-conjugacy classes
1534: $\Pl$ of parabolic $\QQ$-subgroups and we have
1535: \begin{equation*}
1536: \Xbar= \coprod_{P\in \Pl} Y_P\ ,
1537: \end{equation*}
1538: where recall $Y_P=\G_P\back e_P$.  The
1539: bijection $P\mapsto Y_P$ from $\Pl$ to the strata of $\Xbar$ is an
1540: isomorphism of partially ordered sets, where $Y_P\le Y_Q$ if and only if
1541: $Y_P\subseteq \cl{Y_Q}$.  Note that $\cl{Y_Q}$ may be identified with the
1542: Borel-Serre compactification $\Ybar_Q$ of $Y_Q$.
1543: 
1544: \subsection{Local structure of $\Xbar$}
1545: Set $\bar{\mathscr A}_P^G = \mathscr A_P^G \times_{A_P^G} \Abar_P^G$.  The
1546: factorization \eqref{eqnTrivializationPrincipalAPBundle} extends to $D(P)=
1547: \bar{\mathscr A}_P^G \times e_P$.  It follows from reduction theory that
1548: near a point of the stratum $e_P$ of $\Dbar$, the equivalence relation
1549: induced by $\G$ is that of $\G_P$.  Thus $\Xbar$ is locally homeomorphic
1550: along $Y_P$ to $\bar{\mathscr A}_P^G \times Y_P$ (where we embed $Y_P
1551: \subseteq \bar{\mathscr A}_P^G \times Y_P$ as $\{o_P\}\times Y_P$).
1552: Consequently if $\G$ is torsion-free then $\Xbar$ is a smooth
1553: manifold-with-corners.  In general, since any arithmetic group has a
1554: torsion-free subgroup of finite index there may be finite quotient
1555: singularities and $\Xbar$ is a ``$V$-manifold-with-corners''.
1556: 
1557: The description of a neighborhood of the entire closed stratum $\Ybar_P$ is
1558: much more subtle; see \cite[\S8]{refnSaperTilings}.
1559:  
1560: \subsection{Reductive Borel-Serre compactification}
1561: \label{ssectReductiveBorelSerre}
1562: Let $\Xhat$ be the stratified \emph{set}
1563: \begin{equation*}
1564: \Xhat= \coprod_{P\in \Pl} X_P
1565: \end{equation*}
1566: and define a stratified map $\nil\colon \Xbar \to \Xhat$ by setting
1567: $\nil|_{Y_P}$ to be the flat nilmanifold fibration $Y_P\to X_P$ with fiber
1568: $\mathscr N_P(\RR)'$.  We give $\Xhat$ the quotient topology induced by
1569: $\nil$; this is the {\itshape reductive Borel-Serre compactification\/} of
1570: $X$.  (Note that $X=X_G=Y_G$.)  Again $P\mapsto X_P$ is an isomorphism of
1571: $\Pl$ with the partially ordered set of strata of $\Xhat$.  The closure
1572: $\cl{X_Q}$ is identified with the reductive Borel-Serre compactification
1573: $\Xhat_Q$ of $X_Q$.
1574: 
1575: We denote the inclusion map of a stratum $X_P$ by $\i_P\colon
1576: X_P\hookrightarrow \Xhat$ and the inclusion of its closure by
1577: $\ihat_P\colon \Xhat_P \hookrightarrow \Xhat$.  Let $U_P = \coprod_{R\ge P}
1578: X_P$ be the open star neighborhood of a stratum $X_P$; we denote the
1579: inclusion of the deleted star neighborhood by $\j_P\colon U_P\setminus X_P
1580: \hookrightarrow \Xhat$ and the inclusion of the complement of $\Xhat_P$ by
1581: $\jhat_P\colon \Xhat\setminus \Xhat_P \hookrightarrow \Xhat$.
1582: 
1583: \subsection{Local structure of $\Xhat$}
1584: Note that the identification $A_P^G\cong \mathscr A_P^G$ from
1585: \S\ref{ssectGeodesicRetraction} (which depends on the choice of a
1586: basepoint) extends to an identification $\bar{\mathscr A}_P^G\cong
1587: \Abar_P^G$.  The stratification \eqref{eqnAbarStratification} of
1588: $\Abar_P^G$ thus induces one of $\bar{\mathscr A}_P^G$ and this
1589: stratification is independent of the basepoint.
1590: 
1591: Set $\mathscr Z_P = (\bar{\mathscr A}_P^G \times \mathscr N_P(\RR)')/{\sim}$,
1592: where $( a,\G_{N_P} n_1) \sim ( a,\G_{N_P} n_2)$ if and only if $n_2= u
1593: n_1$ for $u\in N_Q(\RR)$ and $a$ belongs to the $Q$-stratum.  The action of
1594: $\G_{L_P}$ on $\mathscr N_P(\RR)'$ induces an action on $\mathscr Z_P$.
1595: Then $\Xhat$ is locally homeomorphic along $X_P$ to the bundle
1596: $D_P\times_{\G_{L_P}} \mathscr Z_P$.
1597: 
1598: Note that if $\G$ is neat then the strata $X_P$ are smooth, however in
1599: general they are only $V$-manifolds.  Also a description of a neighborhood
1600: of the entire closed stratum $\Xhat_P$ may easily be deduced from
1601: \cite[\S8]{refnSaperTilings}.
1602: 
1603: \section{$\L$-modules}
1604: \label{sectLsheaves}
1605: 
1606: \subsection{Notation}
1607: \label{ssectNotationLsheaves}
1608: Let $W\subseteq \Xhat$ be a union of strata, that is, $W= \bigcup_{P\in
1609: \Pl(W)} X_P$ for some subset $\Pl(W)\subseteq \Pl$.  We say $W$ is an
1610: {\itshape admissible\/} space if it is locally closed, or equivalently if
1611: $P$, $Q\in\Pl(W)$ imply that $[P,Q]\subseteq \Pl(W)$.  We will reuse the
1612: notations $\i_P$, $\ihat_P$, $\j_P$, and $\jhat_P$ from
1613: \S\ref{ssectReductiveBorelSerre} in order to denote the analogous
1614: inclusions into $W$.
1615: 
1616: \subsection{}
1617: \label{ssectLstructure}
1618: Let $W$ be an admissible space.  Let $\L_W$ denote the system
1619: consisting of
1620: \begin{equation*}
1621: \begin{cases}
1622: \text{the reductive algebraic groups $L_P$}
1623: &\text{for all $P\in\Pl(W)$, and}\\
1624: \text{the functors }H(\n_P^Q;\cdot)\colon \R(L_Q) \longrightarrow
1625: \Graded(L_P) &\text{for all $P\le Q\in \Pl(W)$.}
1626: \end{cases}
1627: \end{equation*}
1628: Note that the functors $H(\n_P^Q;\cdot)$ are degree nondecreasing and there
1629: are natural isomorphisms (in view of \eqref{eqnKostantDegeneration})
1630: \begin{equation}
1631: H(\n_P^R;\cdot) \cong H(\n_P^Q;H(\n_Q^R;\cdot))\qquad\text{for all
1632: $P\le Q \le R$.} \label{eqnLinkComposition}
1633: \end{equation}
1634: 
1635: \subsection{The category of $\L$-modules}
1636: \label{ssectLsheaves}
1637: Let $\L=\L_W$ for a fixed admissible space $W$.  An {\itshape $\L$-module\/}
1638: $\M=(E_\cdot,f_{\cdot\cdot})$ consists of
1639: \begin{equation*}
1640: \begin{cases}
1641: \text{an object }E_P \text{ of }\Graded(L_P)
1642: &\text{for all $P\in\Pl(W)$, and}\\
1643: \text{degree 1 morphisms }f_{PQ}\colon H(\n_P^Q;E_Q)\xrightarrow{[1]} E_P &
1644: \text{for all $P\le Q\in \Pl(W)$}
1645: \end{cases}
1646: \end{equation*}
1647: satisfying the condition that
1648: \begin{equation}
1649: \sum_{Q\in [P,R]} f_{PQ}\circ H(\n_P^Q;f_{QR}) = 0 \qquad\text{for all
1650: $P\le R\in\Pl(W)$.} 
1651: \label{eqnLsheafCondition}
1652: \end{equation}
1653: (Equation \eqref{eqnLinkComposition} shows this formula makes sense.)
1654: 
1655: Note that \eqref{eqnLsheafCondition} implies in particular  that $(E_P,f_{PP})$
1656: is a complex which we denote $\i_P^!\M$.  (See
1657: \S\ref{ssectFunctorsOnLsheaves} below for a general definition of $k^!$.)
1658: 
1659: An {\itshape $\L$-morphism\/} $\M\to \M'$ is a family
1660: $\phi=(\phi_{\cdot\cdot})$, where
1661: \begin{equation*}
1662: \phi_{PQ}\colon  H(\n_P^Q;E_Q)\to E'_P \qquad  \text{for all $P\le
1663: Q\in\Pl(W)$},
1664: \end{equation*}
1665: such that the following condition is satisfied:
1666: \begin{equation}
1667: \sum_{Q\in [P, R]} \phi_{PQ}\circ
1668: H(\n_P^Q;f_{QR}) = \sum_{Q\in [P,R]} f'_{PQ}\circ H(\n_P^Q;\phi_{QR})
1669: \qquad\text{for all $P\le R\in\Pl(W)$.}
1670: \label{eqnMorphismLsheafCondition}
1671: \end{equation}
1672: The composition of two $\L$-morphisms $\phi\colon \M\to \M'$ and
1673: $\phi'\colon \M'\to \M''$ is defined by
1674: \begin{equation}
1675: (\phi'\circ\phi)_{PR} = \sum_{Q\in [P,R]}
1676: \phi'_{PQ} \circ H(\n_P^Q;\phi_{QR}) \qquad\text{for all $P\le R\in\Pl(W)$.}
1677: \label{eqnCompositionLMorphism}
1678: \end{equation}
1679: 
1680: The {\itshape category of $\L$-modules\/} $\R(\L)$ has for objects the
1681: $\L$-modules and for morphisms the $\L$-morphisms.  The full subcategory
1682: $\R^b(\L)$ consists of $\L$-modules $\M$ for which all $E_P$ lie in
1683: $\Graded^b(L_P)$.  Note that if $W=X_P$ consists of a single stratum, then
1684: $\R(\L_{X_P})=\Complex(L_P)$, however in general $\R(\L)$ is not the
1685: category of complexes $\Complex(\Cat)$ for some category $\Cat$.
1686: 
1687: \subsection{Standard functors on $\L$-modules}
1688: \label{ssectFunctorsOnLsheaves}
1689: Given an inclusion $k\colon Z\hookrightarrow W$ of admissible spaces one can
1690: define functors $k^*$, $k_*$, $k^!$, and $k_!$ between the respective
1691: categories of $\L$-modules that are motivated by the usual functors on the
1692: derived category of sheaves.
1693: 
1694: The functor $k^!$ associates to an $\L_W$-module $\M$ an $\L_Z$-module with
1695: the same data but with $P$ and $Q$ restricted to belong to $\Pl(Z)$.
1696: Condition \eqref{eqnLsheafCondition} continues to hold since $Z$ is locally
1697: closed.  One special case is $\i_P^!\M = (E_P,f_{PP})$, the \emph{local
1698: cohomology complex at $P$ with supports}.  The functor $k_*$ associates to
1699: an $\L_Z$-module $\M$ an $\L_W$-module with the same data but extended by
1700: $E_P=0$ and $f_{PQ}=0$ if one of the subscripts is outside of $\Pl(Z)$.
1701: 
1702: We only define $k^*$ in the cases we need, namely when
1703: $Z$ is open in $W$ (in which case $k^*=k^!$ has been
1704: defined above) and when $Z$ has a unique maximal face.
1705: Suppose $X_T$ is a maximal face of $Z$; that is,
1706: $T\in\Pl(Z)$ and $Z\subseteq \Xhat_T$.  For
1707: $\M=(E_\cdot,f_{\cdot\cdot})$ an $\L_W$-module, define $k^*\M =
1708: (E'_\cdot,f'_{\cdot\cdot})$ by
1709: \begin{equation*}
1710: \left\{
1711: \begin{aligned}
1712: E'_P&= \bigoplus_{R \in [P,(P,T)]} H(\n_P^R;E_R), \\
1713: f'_{PQ}&= \sum_{\substack{R\cap T =P \\ S\cap T=Q \\ R\le S}} H(\n_P^R;f_{RS}),
1714: \end{aligned}
1715: \right.
1716: \end{equation*}
1717: for all $P\le Q$ with $P$, $Q\in\Pl(Z)$.  (Recall that $(P,T)$ denotes the
1718: parabolic $\QQ$-subgroup opposite to $T$ relative to $P$.)  The reader can
1719: verify this is an $\L_Z$-module.  As a special case, the {\itshape local
1720: cohomology complex at $P$\/} is
1721: \begin{equation*}
1722: \i_P^*\M = \biggl( \bigoplus_{P\le R} H(\n_P^R;E_R), 
1723: \sum_{R\in [P,S]} H(\n_P^R;f_{RS}) \biggr).
1724: \end{equation*}
1725: It should always be understood in these and similar formulas that $R$ and
1726: $S$ are restricted to belong to $\Pl(W)$.
1727: 
1728: We only define $k_!$ in the case we need, namely when
1729: $Z$ is closed in $W$, in which case $k_!=k_*$ has been
1730: defined above.
1731: 
1732: \begin{prop*}
1733: Assume in the following that $j$ and $k$ are inclusions of admissible
1734: spaces and that the indicated functors have been defined above.
1735: \begin{enumerate}
1736: \item There are identities\/\textup:
1737: \begin{alignat*}{2}
1738: j^!\circ k^! &= (k\circ j)^!\ , &\qquad\qquad k_*\circ j_* &= (k\circ j)_*
1739: \ ,\\
1740: j^*\circ k^* &= (k\circ j)^*\ , &\qquad\qquad k_!\circ j_! &= (k\circ j)_!
1741: \ .
1742: \end{alignat*}
1743: \label{itemFunctorsComposition}
1744: \item There are adjoint relations\/\textup:
1745: \begin{equation}
1746: \begin{aligned}
1747: \Hom_{\R(\L_W)}(\M,k_*\M') &\cong
1748:      \Hom_{\R(\L_Z)}(k^*\M,\M') , \\
1749: \Hom_{\R(\L_W)}(\M,k^!\M') &\cong
1750:      \Hom_{\R(\L_Z)}(k_!\M,\M').
1751: \end{aligned}\label{eqnAdjointRelations}
1752: \end{equation}
1753: \label{itemFunctorsAdjoint}
1754: \item Consider the commutative diagram of inclusions of admissible
1755: spaces,
1756: \begin{equation*}
1757: \begin{CD}
1758: Z @>k_2>> Z_2 \\
1759: @V j_1 VV     @V j_2 VV \\
1760: Z_1 @>k_1>> W \rlap{\ .}
1761: \end{CD}
1762: \end{equation*}
1763: where $Z=Z_1\cap Z_2$.  Then
1764: \begin{equation}
1765: k_1^!\circ  j_{2*} = j_{1*}\circ k_2^!\ .
1766: \end{equation}
1767: Now assume that $Z_1$ has a unique maximal face $X_T$, that $Z$ has a
1768: unique maximal face $X_{T_0}$, and furthermore that $Z_1$ and $Z_2$
1769: intersect ``transversely'' in the sense that
1770: \begin{equation}
1771: U\cap Z_2 = U\cap W\cap\Xhat_{(T_0,T)},\label{eqnTransverse}
1772: \end{equation}
1773: where $U = \bigcup_{P\in\Pl(Z)}\bigcup_{Q\ge P} X_Q$ is the star
1774: neighborhood of $Z$.  Then
1775: \begin{equation*}
1776: j_1^!\circ k_1^* = k_2^* \circ j_2^!.
1777: \end{equation*}
1778: \label{itemFunctorsSquare}
1779: \end{enumerate}
1780: \end{prop*}
1781: \begin{proof}
1782: Parts \itemref{itemFunctorsComposition} and \itemref{itemFunctorsAdjoint}
1783: and the first part of \itemref{itemFunctorsSquare} are left to the reader.
1784: For the rest of \itemref{itemFunctorsSquare}, note that $E'_P = \bigoplus_Q
1785: H(\n_P^Q;E_Q)$ for both $j_1^! k_1^*\M$ and $k_2^* j_2^!\M$, where in the
1786: first case the sum is over $Q\in [P,(P,T)]\cap\Pl(W)$, while in the second
1787: case the sum is over $Q\in [P,(P,T_0)]\cap\Pl(Z_2)$.  The sums are equal by
1788: \eqref{eqnTransverse} since $(P,T)= (P,T_0)\cap (T_0,T)$.
1789: \end{proof}
1790: 
1791: Elsewhere we will define $k^*$ and $k_!$ more generally and prove an
1792: analogue of the proposition in a suitable homotopy category of $\L$-modules
1793: (see \S\ref{ssectHomotopyLModules}).
1794: 
1795: \subsection{Pullback of an $\L_W$-module to a fiber}
1796: \label{ssectPullbackToFiber}
1797: There is one other functor we will need to consider beginning in
1798: \S\ref{sectEqualRankMicropurityNEW}.  Suppose $G$ has a connected normal
1799: $\QQ$-subgroup which we denote $G_{\l}$.  Let $\tilde G_h$ be a
1800: complementary connected normal $\QQ$-subgroup so that $G=\tilde G_h G_\l$
1801: is an almost direct product.  Set $G_{h} = G/G_{\l}=\tilde
1802: G_h/(G_\l\cap \tilde G_h)$ and let $\G_{G_{\l}}=\G\cap G_{\l}$ and
1803: $\G_{G_{h}}=\G/ \G_{G_{\l}}$ be the induced arithmetic subgroups.
1804: Parabolic $\QQ$-subgroups of $G$ correspond to pairs of a parabolic
1805: $\QQ$-subgroup of $G_\l$ and a parabolic $\QQ$-subgroup of $G_h$: given
1806: $P\subseteq G$ we associate $P_\l=P\cap G_\l$ and $P_h=P/P_\l$ and
1807: conversely given $P_\l$ and $P_h$ we associate $P=\tilde P_h P_\l$
1808: (where $\tilde P_h\subseteq \tilde G_h$ is a lift of $P_h$).
1809: 
1810: The geometric picture is as follows.  There is a factorization of symmetric
1811: spaces $D = D_{h}\times D_{\l}$ and a flat bundle of arithmetic quotients
1812: $X \to X_{h}$ with fibers $X_\l$.  (The action of $\G_{G_h}$ is induced
1813: from the action of the finite quotient $\G/\G_\l(\G\cap \tilde G_h)$, so
1814: this bundle becomes trivial over a finite cover of $X_h$.)  Let
1815: $X(G_\l)\subseteq \Xhat$ be the partial compactification of $X$ obtained by
1816: replacing the fibers by $\Xhat_\l$.  Specifically, note that the map
1817: $P\mapsto P_h$ induces a surjection $\Pl(\Xhat) \to \Pl(\Xhat_h)$.  Then
1818: set $X(G_\l)$ to be the union of strata $X_P$ for all $P\in\Pl(\Xhat)$
1819: satisfying $P_h=G_{h}$.  The above fibration extends to a fibration
1820: \begin{equation*}
1821: \pi\colon  X(G_\l) \longrightarrow X_{h}
1822: \end{equation*}
1823: with fibers $\Xhat_{\l}$.  Let
1824: $\ihat_{G,\l}\colon \Xhat_{\l}\hookrightarrow X(G_\l)$ be the inclusion of a
1825: generic fiber.
1826: 
1827: The map $P_\l\mapsto P=\tilde G_h P_\l$ induces a surjection
1828: $\Pl(\Xhat_\l)\to \Pl(X(G_\l))$.  (It may not be injective if the
1829: finite group $\G/\G_\l(\G\cap \tilde G_h)$ is not trivial.)  Define the
1830: connected normal $\QQ$-subgroup $L_{P,\l} = P_\l/N_P\subseteq L_P$.
1831: Let $\M=(E_\cdot,f_{\cdot\cdot})$ be an $\L_{X(G_\l)}$-module.  Define an
1832: $\L_{\Xhat_{\l}}$-module $\ihat_{G,\l}^*\M =
1833: (E'_\cdot,f'_{\cdot\cdot})$ by
1834: \begin{equation*}
1835: \left\{
1836: \begin{aligned}
1837: E'_{P_\l}&= \Res_{L_{P,\l}}^{L_P} E_P, \\
1838: f'_{P_\l Q_\l}&= \Res_{L_{P,\l}}^{L_P} f_{PQ},
1839: \end{aligned}
1840: \right.
1841: \end{equation*}
1842: for all $P_\l\le Q_\l\in\Pl(\Xhat_\l)$.  (Here
1843: $\Res_{L_{P,\l}}^{L_P}$ denotes the restriction of a representation
1844: of $L_P$ to a representation of $L_{P,\l}$.)  This makes sense
1845: since there is a natural isomorphism
1846: \begin{equation*}
1847: \Res_{L_{P,\l}}^{L_P} H(\n_P^Q;E_Q) \cong
1848: H(\n_{P_\l}^{Q_\l};\Res_{L_{Q,\l}}^{L_Q} E_Q)\ .
1849: \end{equation*}
1850: Similarly define $\ihat_{G,\l}^!\M$ except that in this case
1851: $E'_{P_\l} = \Res_{L_{P,\l}}^{L_P} E_P[-\dim D_h]$.
1852: 
1853: More generally for $R\in\Pl$ we may define $\ihat_{R,\l}^*$ and
1854: $\ihat_{R,\l}^!$ when we are given a connected normal $\QQ$-subgroup
1855: $L_{R,\l}$ of $L_R$.  In this case we have a fibration
1856: $X_R(L_{R,\l})\to X_{R,h}$ with fiber $\Xhat_{R,\l}$ and a surjection
1857: $\Pl(\Xhat_{R,\l})\to \Pl(X_R(L_{R,\l}))$ given by $P_\l\mapsto P$
1858: where $P/N_R = \widetilde{L_{R,h}} P_\l $.  We note the following
1859: identities for $P_\l\in\Pl(\Xhat_{R,\l})$:
1860: \begin{equation}
1861: \begin{alignedat}{2}
1862: \i_{P_\l}^* \circ \ihat_{R,\l}^* &= \i_{P_\l}^* \circ \ihat_{P,\l}^*
1863: \circ \i_P^*\ , &\qquad\qquad
1864: \i_{P_\l}^* \circ \ihat_{R,\l}^! &= \i_{P_\l}^* \circ \ihat_{P,\l}^!
1865: \circ \i_P^*\ , \\
1866: \i_{P_\l}^! \circ \ihat_{R,\l}^* &= \i_{P_\l}^* \circ \ihat_{P,\l}^*
1867: \circ \i_P^!\ , &\qquad\qquad
1868: \i_{P_\l}^! \circ \ihat_{R,\l}^! &= \i_{P_\l}^* \circ \ihat_{P,\l}^!
1869: \circ \i_P^!\ .
1870: \end{alignedat}
1871: \label{eqnPullbackToFiberFormulas}
1872: \end{equation}
1873: (The initial $\i_{P_\l}^*$ on the right-hand side of these equalities
1874: would not be needed if $\G/\G_\l(\G\cap \tilde G_h)$ were trivial.)
1875: 
1876: \subsection{Local cohomology with supports}
1877: \label{ssectLocalCohomologyWithSupports}
1878: Let $\M$ be an $\L_W$-module, where $W$ is an admissible space.  It is easy
1879: to verify that there is a short exact sequence
1880: \begin{equation}
1881: 0 \to \i_Q^!\M \to \i_Q^*\M \to  \i_Q^*\j_{Q*}\j_Q^*\M \to 0
1882: \end{equation}
1883: relating the two types of local cohomology complexes at $Q$.  There is also
1884: a short exact sequence
1885: \begin{equation}
1886: 0 \to \ihat_Q^!\M \to \ihat_Q^*\M \to \ihat_Q^*\jhat_{Q*}\jhat_Q^*\M \to 0\
1887: .
1888: \label{eqnStandardShortExactEquenceForClosedStratum}
1889: \end{equation}
1890: 
1891: More generally, for $P\le Q\in \Pl(W)$ define the {\itshape local
1892: cohomology complex of $\M$ at $P$ supported on $Q$\/} to be
1893: \begin{equation*}
1894: \i_P^*\ihat_Q^!\M = \biggl( \bigoplus_{R \in [P,Q]} H(\n_P^R;E_R), 
1895: \sum_{R\le S \in [P,Q]} H(\n_P^R;f_{RS}) \biggr)\ ;
1896: \end{equation*}
1897: this will play an essential role in the definition of micro-support of $\M$
1898: later in \S\ref{sectMicroSupport}.  If $Q\le Q'$ there is a short exact
1899: sequence comparing $\i_P^*\ihat_Q^!\M$ and $\i_P^*\ihat_{Q'}^!\M$, namely
1900: \begin{equation}
1901: 0\to \i_P^* \ihat_{Q}^!  \M \to \i_P^* \ihat_{Q'}^! \M \to \i_P^*
1902: \jhat_{Q*}\jhat_Q^*\ihat_{Q'}^! \M \to 0 \ ;
1903: \label{eqnShortCompareLocalCohomologyWithSupports}
1904: \end{equation}
1905: this follows from \eqref{eqnStandardShortExactEquenceForClosedStratum} by
1906: replacing $\M$ with $\ihat_{Q'}^!\M$ and applying $\i_P^*$.  The
1907: corresponding long exact sequence is
1908: \begin{equation}
1909: \cdots \longrightarrow
1910: H^i(\i_P^* \ihat_Q^!  \M) \longrightarrow
1911: H^i(\i_P^* \ihat_{Q'}^! \M) \longrightarrow
1912: H^i(\i_P^* \jhat_{Q*}\jhat_Q^*\ihat_{Q'}^! \M) \longrightarrow \cdots\ .
1913: \label{eqnLongCompareLocalCohomologyWithSupports}
1914: \end{equation}
1915: 
1916: We wish to study $H(\i_P^* \jhat_{Q*}\jhat_Q^*\ihat_{Q'}^! \M)$.  Set
1917: $P'= (P,Q)\cap Q'$ so that $P=P'\cap Q$ and $Q'=P'\vee Q$:
1918: \begin{equation}
1919: \vcenter{\xymatrix @ur @M=1pt @R=1.5pc @C=1pc {
1920: {P'} \ar@{-}[r] \ar@{-}[d] & {Q'} \ar@{-}[d] \\
1921: {P\rlap{\qquad .}} \ar@{-}[r] & {Q}
1922: }}
1923: \end{equation}
1924: When $\#\D_P^{P'}=1$ the last term of
1925: \eqref{eqnShortCompareLocalCohomologyWithSupports} is equal to
1926: \begin{equation*}
1927: \i_P^* \i_{P'*}\i_{P'}^*\ihat_{Q'}^! \M =
1928: H(\n_P^{P'}; \i_{P'}^*\ihat_{Q'}^! \M)\ ;
1929: \end{equation*}
1930: since the functor $H(\n_P^{\tilde P};\cdot)$ is exact it commutes with
1931: taking cohomology and so
1932: \begin{equation}
1933: H(\i_P^* \i_{P'*}\i_{P'}^*\ihat_{Q'}^! \M) =
1934: H(\n_P^{P'}; H(\i_{P'}^*\ihat_{Q'}^! \M))\ .
1935: \label{eqnPushForwardType}
1936: \end{equation}
1937: In general we have the
1938: 
1939: \begin{lem}
1940: \label{ssectRelativeLocalCohomologySupportsSS}
1941: Given $P\le Q\le Q'$, set $P'= (P,Q)\cap Q'$.  There are two spectral
1942: sequences abutting to $H(\i_P^* \jhat_{Q*}\jhat_Q^*\ihat_{Q'}^!  \M)$.  The
1943: first \textup(the Fary spectral sequence\textup) has
1944: \begin{equation}
1945: E_1^{-p,\cdot} =
1946: \bigoplus_{\substack{P < \tilde P \le P' \\ \#\D_P^{\tilde P} = p}}
1947: H(\n_P^{\tilde P};H(\i_{\tilde P}^*\ihat_{\tilde Q}^!\M))[-p] \ ,
1948: \label{eqnFarySpectralSequence}
1949: \end{equation}
1950: where $\tilde Q = \tilde P\vee Q$.  The second \textup(the Mayer-Vietoris
1951: spectral sequence\textup) has
1952: \begin{equation}
1953: E_1^{p,\cdot} =
1954: \bigoplus_{\substack{P < \tilde P \le P' \\ \#\D_P^{\tilde P} = p+1}}
1955: H(\n_P^{\tilde P};H(\i_{\tilde P}^*\ihat_{Q'}^!\M)) \ .
1956: \label{eqnMayerVietorisSpectralSequence}
1957: \end{equation}
1958: \end{lem}
1959: \begin{proof}
1960: Any $R\in [P,Q']$ belongs to $[\tilde P,\tilde Q]$ for a unique $\tilde
1961: P\in [P,P']$ (namely $\tilde P=R\cap P'$).  The situation is represented by
1962: the following diagram, in which each parallelogram has the analogous
1963: meaning to \eqref{eqnParallelogram}:
1964: \begin{equation}
1965: \vcenter{\xymatrix @ur @M=1pt @R=1.5pc @C=1pc {
1966: {P'} \ar@{-}[rr] \ar@{-}[d] & {} & {Q'} \ar@{-}[d] \\
1967: {\tilde P} \ar@{-}[r] \ar@{-}[d] & {R} \ar@{-}[r] & {\tilde Q}
1968: \ar@{-}[d] \\
1969: {P\rlap{\qquad\qquad .}} \ar@{-}[rr] & {} & {Q}
1970: }} \label{eqnParallelogramWithPtilde}
1971: \end{equation}
1972: (Of course this figure is not meant to suggest that the parabolic
1973: $\QQ$-subgroups lying between $P$ and $P'$ are linearly ordered.)
1974: 
1975: For the first spectral sequence, decompose $\i_P^*
1976: \jhat_{Q*}\jhat_Q^*\ihat_{Q'}^! \M$ (as an {\itshape $L_P$-module\/}, not
1977: as a complex) as
1978: \begin{equation}
1979: \bigoplus_{P<\tilde P\le P'}\i_P^* \i_{\tilde P*}\i_{\tilde P}^*
1980:      \ihat_{\tilde Q}^! \M
1981: =\bigoplus_{P<\tilde P\le P'} H(\n_P^{\tilde P};\i_{\tilde P}^*
1982:      \ihat_{\tilde Q}^! \M) \ .
1983: \label{eqnTildePDecomposition}
1984: \end{equation}
1985: The corresponding sum in which $\tilde P$ is restricted to those with
1986: $\#\D_P^{\tilde P} \le p$ is a subcomplex and defines an increasing
1987: filtration on $\i_P^* \jhat_{Q*}\jhat_Q^*\ihat_{Q'}^! \M$ as $p$ varies.
1988: The associated graded complex is exactly equal to
1989: \eqref{eqnTildePDecomposition}.  In view of \eqref{eqnPushForwardType}, the
1990: $E_1$ term of the resulting spectral sequence has the desired form,
1991: \eqref{eqnFarySpectralSequence}.
1992: 
1993: For the Mayer-Vietoris spectral sequence, consider the long exact sequence
1994: \begin{equation*}
1995: 0\to \i_P^* \jhat_{Q*}\jhat_Q^*\ihat_{Q'}^! \M \to 
1996: \bigoplus_{\#\D_P^{\tilde P}=1} \i_P^*\i_{{\tilde P}*}\i_{\tilde P}^*
1997:          \ihat_{Q'}^!\M \to
1998: \bigoplus_{\#\D_P^{\tilde P}=2} \i_P^*\i_{{\tilde P}*}\i_{\tilde P}^*
1999:          \ihat_{Q'}^!\M \to
2000: \cdots
2001: \end{equation*}
2002: in which $\tilde P$ satisfies $P< \tilde P\le P'$.  Thus $H(\i_P^*
2003: \jhat_{Q*}\jhat_Q^*\ihat_{Q'}^! \M)$ is the cohomology of a double complex;
2004: the spectral sequence \eqref{eqnMayerVietorisSpectralSequence} corresponds
2005: to the ``first filtration'' $\#\D_P^{\tilde P} \ge p+1$.
2006: \end{proof}
2007: 
2008: \subsection{}
2009: \label{ssectMappingCone}
2010: The {\itshape mapping cone\/} of an $\L$-morphism
2011: $\phi\colon \M\to \M'$ is defined by
2012: \begin{equation*}
2013: M(\phi)= \left(E_\cdot[1]\oplus E'_\cdot\, , 
2014: \left(\begin{matrix} -f_{\cdot\cdot} & 0 \\
2015:             \phi_{\cdot\cdot} & f'_{\cdot\cdot} \end{matrix}\right)\right).
2016: \end{equation*}
2017: This is an $\L$-module; condition \eqref{eqnLsheafCondition} follows from
2018: the corresponding condition for $\M$ and $\M'$ and from
2019: \eqref{eqnMorphismLsheafCondition}.  The mapping cone defines a functor from the
2020: category of diagrams $\phi\colon \M\to \M'$ to $\R(\L)$. Define the {\itshape shift\/}
2021: of an $\L$-module by $\M[1]= (E_\cdot[1],-f_{\cdot\cdot})$.  There are
2022: natural $\L$-morphisms $\al(\phi)\colon \M'\to M(\phi)$ and $\b(\phi)\colon M(\phi)\to
2023: \M[1]$ given by
2024: \begin{alignat*}{2}
2025: \al(\phi)_{PP} &= \begin{pmatrix} 0 \\ \id_{E'_P}\end{pmatrix} &&\qquad\text{for all
2026: $P$}, \\
2027: \b(\phi)_{PP} &= \begin{pmatrix} \id_{E_P[1]} & 0\end{pmatrix} &&\qquad\text{for all
2028: $P$}.
2029: \end{alignat*}
2030: (Set $\al(\phi)_{PQ}=0$ for $P\lneq Q$ and similarly for $\b(\phi)$.)
2031: 
2032: \subsection{}
2033: \label{ssectHomotopyLModules}
2034: One may define a notion of homotopy for $\L$-morphisms and set $\K(\L)$ to
2035: be the category with $\L$-modules as objects and homotopy classes of
2036: $\L$-morphisms as morphisms.  With the definition of mapping cone above,
2037: $\K(\L)$ may be given the structure of a {\itshape triangulated category\/}
2038: \cite[\S\S1.4,~ 1.5]{refnKashiwaraSchapira}.  A {\it quasi-isomorphism\/}
2039: of $\L$-modules is an $\L$-morphism that induces quasi-isomorphisms on
2040: local cohomology complexes.  Since $\R(L_P)$ is a semisimple abelian
2041: category for all $P$ one can show that any quasi-isomorphism of $\L$-modules
2042: has a homotopy inverse.  Thus it is reasonable to notate $\K(\L)$ also as
2043: $\Derived(\L)$.  We will discuss this in more detail elsewhere.
2044: 
2045: \subsection{Generalizations}
2046: \label{ssectLModuleGeneralizations}
2047: The concept of $\L$-modules may be greatly generalized.  For example, in
2048: \S\ref{ssectLstructure} one could start with a general family of reductive
2049: $\QQ$-groups $(L_\sigma)_\sigma$ indexed by a partially ordered set and
2050: replace $H(\n_P^Q;\cdot)$ by some ``link cohomology'' functors
2051: $\M_{\sigma\tau}\colon \R(L_\tau) \longrightarrow \Graded(L_\sigma)$ satisfying
2052: $\M_\sigma^\upsilon \cong \M_\sigma^\tau\circ \M_\tau^\upsilon$.  One
2053: particularly simple situation will be introduced in
2054: \S\ref{ssectLModulesAPTimesXP} in which all the groups $L_\sigma$ will be
2055: identical.  One could also drop the condition that the groups are reductive
2056: algebraic and replace $\R(L_\sigma)$ by a suitable category of
2057: representations.  We will not consider such generalizations further
2058: here.
2059: 
2060: \section{Realization of $\L$-modules}
2061: \label{sectRealizationLModules}
2062: For an admissible space $W$ let $\X$ denote the stratification $W =
2063: \coprod_{P\in\Pl(W)} X_P$.  Recall that $\Complex_\X(W)$ denotes the
2064: category of constructible complexes of sheaves on $W$ and that
2065: $\Derived_\X(W)$ denotes the corresponding derived category.  Note that a
2066: regular $L_P$-module $E_P$ induces by restriction a $\G_{L_P}$-module and
2067: hence a locally constant sheaf $\EE_P=E_P \times_{\G_P} D_P$ on $X_P$.
2068: Thus we obtain a functor $\R(L_P)\to \Derived_\X(X_P)$.  In this section we
2069: generalize this to $\L$-modules in order to obtain functors $\Sheaf_W\colon
2070: \R(\L_W)\to \Derived_{\X}(W)$.
2071: 
2072: \begin{thm}
2073: \label{ssectRealizationLModules}
2074: \ \par
2075: \begin{list}{\labelenumi}{\usecounter{enumi}\def\makelabel#1{\hss\llap{\upshape#1}}\setlength{\leftmargin}{0pt}\setlength{\itemindent}{7ex}}
2076: \item
2077: There exists a family of functors
2078: \begin{equation*}
2079: \A_P\colon \R(L_P) \to \Complex_{\X}(\Xhat_P) \qquad\text{for all
2080: $P\in\Pl$}
2081: \end{equation*}
2082: and natural morphisms  for all $P\le Q\in\Pl$,
2083: \begin{equation}
2084: k_{PQ}\colon \ihat_{Q*}\A_Q(E_Q)\longrightarrow \ihat_{P*}\A_P(H(\n_P^Q;E_Q))
2085: \qquad\text{for all $E_Q\in\R(L_Q)$.}
2086: \label{eqnRealizationPQ}
2087: \end{equation}
2088: Each $\A_P$ is an incarnation of the functor $E_P\mapsto \i_{P*}\EE_P$,
2089: $R(L_P) \to \Derived_{\X}(\Xhat_P)$, and $\ihat_P^*(k_{PQ})$ becomes a
2090: natural isomorphism in $\Derived_\X(\Xhat_P)$.
2091: \label{itemExistenceOfBasicRealizationFunctors}
2092: \item
2093: A family of functors as in
2094: \itemref{itemExistenceOfBasicRealizationFunctors}
2095: determines functors
2096: \begin{equation*}
2097: \A_W\colon \R(\L_W)\to \Complex_{\X}(W)\qquad \text{for all admissible spaces
2098: $W$, and}
2099: \end{equation*}
2100: corresponding functors $\Sheaf_W$ into $\Derived_{\X}(W)$.  If
2101: $k\colon Z\hookrightarrow W$ is an inclusion of admissible spaces, then $k_*\circ
2102: \Sheaf_Z= \Sheaf_W \circ k_*$ and similarly for the other functors defined
2103: in \S\S\textup{\ref{ssectFunctorsOnLsheaves}} and
2104: \textup{\ref{ssectPullbackToFiber}}.
2105: \label{itemExistenceOfRealizationFunctors}
2106: \item
2107: If two families of functors as in
2108: \itemref{itemExistenceOfBasicRealizationFunctors} are naturally
2109: quasi-isomorphic compatibly with \eqref{eqnRealizationPQ}, then the
2110: corresponding functors $\Sheaf_W$ are naturally isomorphic.
2111: \label{itemUniquenessOfRealizationFunctors}
2112: \end{list}
2113: \end{thm}
2114: 
2115: \begin{rem*}
2116: A functor $\Sheaf_W$ as in the theorem with be called a {\itshape
2117: realization\/} of the category of $\L_W$-modules.  Given a realization
2118: functor and an $\L_W$-module $\M$, the {\itshape cohomology $H(W;\M)$ of
2119: $\M$\/} is defined to be the hypercohomology $H(W;\Sheaf_W(\M))$.  More
2120: generally the cohomology $H(Y;\M)$ of any subset $Y\subseteq W$ with
2121: coefficients in $\M$ is defined to be the hypercohomology of
2122: $\Sheaf_W(\M)|_Y$.  In this paper we will use the incarnation $\A_W(\M)$
2123: constructed below using special differential forms.  This is a complex of
2124: fine sheaves so $H(\Xhat;\M)$ may be computed from the complex of global
2125: sections of $\A_{\Xhat}(\M)$.
2126: \end{rem*}
2127: 
2128: \subsection{Proof of Theorem \ref{ssectRealizationLModules}\itemref{itemExistenceOfRealizationFunctors}\itemref{itemUniquenessOfRealizationFunctors}}
2129: \label{ssectRealizationLsheaves}
2130: We assume that a family of functors $\{\A_P\}_{P\in\Pl}$ exists as in
2131: \itemref{itemExistenceOfBasicRealizationFunctors}.  Let
2132: $\M=(E_\cdot,f_{\cdot\cdot})$ be an $\L_W$-module.  Define
2133: \begin{equation}
2134: \left\{
2135: \begin{aligned}
2136: \A_W(\M) &= \bigoplus_{P\in \Pl(W)} 
2137:                             \ihat_{P*} \A_P(E_P), \\
2138: d_{\A_W(\M)} &= \sum_{P\in \Pl(W)} d_P + \sum_{P\le Q\in \Pl(W)}
2139:                             \A_P(f_{PQ}) \circ k_{PQ} \ ,
2140: \end{aligned}
2141: \right.
2142: \label{eqnRealizationComplex}
2143: \end{equation}
2144: where $d_P$ is the differential of $\A_P(E_P)$.  (Note that in applying
2145: $\A_P$ to $E_P$ we are using the sign convention of
2146: \S\ref{sssectNotationCategory}.)  This is a complex of sheaves.  For
2147: $\phi=(\phi_{\cdot\cdot})$ an $\L_W$-morphism define
2148: \begin{equation*}
2149: \A_W(\phi) = \sum_{P\le Q} \A_P(\phi_{PQ})\circ
2150: k_{PQ} \ .
2151: \end{equation*}
2152: This is the desired functor for
2153: \itemref{itemExistenceOfRealizationFunctors}; we let $\Sheaf_W$ be the
2154: corresponding functor into $\Derived_\X(W)$.  It is easy to verify that
2155: $\Sheaf_W$ commutes with the usual functors defined in
2156: \S\S\ref{ssectFunctorsOnLsheaves} and \ref{ssectDegreeTruncationLmodules}
2157: and that \itemref{itemUniquenessOfRealizationFunctors} holds.
2158: 
2159: \subsection{Special differential forms}
2160: We now prepare for the proof of Theorem
2161: \ref{ssectRealizationLModules}\itemref{itemExistenceOfBasicRealizationFunctors}
2162: which will be at the end of this section.
2163: 
2164: Let $E$ be a regular representation of $G$ and let $\EE$ denote the
2165: corresponding locally constant sheaf on $\Xbar$, or on any stratum of
2166: $\Xbar$.  For any stratum $Y_P$ of $\Xbar$, consider the diagram
2167: \begin{equation*}
2168: \Xbar = \G\back \overline{D} \overset{q}{\longleftarrow} \G_P\back D(P)
2169: \overset{r_P}{\longrightarrow}  Y_P,
2170: \end{equation*}
2171: where $q\colon \G_P\back D(P)\hookrightarrow \G_P\back\overline{D} \to
2172: \G\back\overline{D}$ is the inclusion followed by the quotient and $r_P$ is
2173: the geodesic retraction.  Reduction theory implies that if $y\in Y_P$ there
2174: exists a sufficiently small neighborhood $U\subseteq \Xbar$ of $y$ such
2175: that $q\colon q^{-1}(U)\tildearrow U$ is a diffeomorphism.  An $\EE$-valued
2176: differential form $\o$ on $\Xbar$ is {\itshape locally lifted from the
2177: boundary\/} \cite[\S8]{refnBorelStable} if for all strata $Y_P$ and all
2178: $y\in Y_P$, there exists $U$ as above so that $q^*(\o|_U) = r_P^*\psi$
2179: where $\psi$ is a smooth form on $U\cap Y_P$.  A form on a stratum
2180: $Y_P=\G_P\back e_P$ is called {\it $N_P(\RR)$-invariant\/} if its lift to
2181: $e_P$ is invariant under the action of $N_P(\RR)$; let $\A\inv(Y_P;\EE)$
2182: denote the sheaf of $N_P(\RR)$-invariant $\EE$-valued forms on $Y_P$ and
2183: let $A\inv(Y_P;\EE)$ denote the global sections.
2184: 
2185: Let $\Asp(\Xbar;\EE)\subseteq \A(\Xbar;\EE)$ denote the subcomplex of forms
2186: which are locally lifted from the boundary and whose restriction to each
2187: stratum $Y_P$ is in $\A\inv(Y_P;\EE)$.  The sheaf complex of {\itshape
2188: special differential forms\/} on $\Xhat$ is
2189: \cite[(13.2)]{refnGoreskyHarderMacPherson}
2190: \begin{equation*}
2191: \Asp(\Xhat;\EE) = \nil_{*} \Asp(\Xbar;\EE)\ ,
2192: \end{equation*}
2193: the pushforward under the quotient map $\nil\colon \Xbar\to \Xhat$.  Let
2194: $A\sp(\Xbar;\EE)$ and $A\sp(\Xhat;\EE)$ denote the corresponding complexes
2195: of global sections.
2196: 
2197: \subsection{Restriction to boundary strata}
2198: Recall from \S\ref{ssectNilmanifoldFibrations} that the nilmanifold
2199: fibration $\nil\colon  Y_P \to X_P$ is the flat bundle associated to the
2200: $\G_{L_P}$-space $\mathscr N_P(\RR)'$, and that given a basepoint $x\in D$
2201: there is an isomorphism $n_x'\colon  N_P(\RR)'\tildearrow \mathscr N_P(\RR)'$.
2202: The $\G_{L_P}$-action on $\mathscr N_P(\RR)'$ induces the structure of an
2203: $\G_{L_P}$-module on $A\inv(\mathscr N_P(\RR)';\EE)$, the
2204: $N_P(\RR)$-invariant forms; let $\AA\inv(\mathscr N_P(\RR)';\EE)$ denote%
2205: %
2206: \footnote{In \cite{refnGoreskyHarderMacPherson} this complex is denoted
2207: $\mathbf C^{\boldsymbol \cdot}(N_P;\mathbf E)$.}
2208: %
2209: the associated complex of locally constant sheaves on $X_P$.  The fiber
2210: $\AA\inv(\mathscr N_P(\RR)';\EE)_x$ is the complex of $N_P(\RR)$-invariant
2211: forms on $\nil^{-1}(x)$.
2212: 
2213: A special differential form $\o$ on $\Xhat$ determines an
2214: $N_P(\RR)$-invariant form on $Y_P$ which we denote $\o|_{Y_P}$; this form
2215: can be viewed as a differential form on $X_P$ with values in
2216: $\AA\inv(\mathscr N_P(\RR)';\EE)$ which we denote $\o|_{X_P}$. In fact there
2217: is a natural isomorphism of complexes \cite[12.6,
2218: 13.4(2)]{refnGoreskyHarderMacPherson}
2219: \begin{equation}
2220: \label{eqnRestrictSpecialForms}
2221: \i_P^*\Asp(\Xhat;\EE)  \longtildearrow
2222: \A(X_P;\AA\inv(\mathscr N_P(\RR)';\EE)),
2223: \end{equation}
2224: where the complex on the right is the associated total complex.  This
2225: operation extends to an isomorphism
2226: \begin{equation*}
2227: \ihat_P^*\Asp(\Xhat;\EE)\longtildearrow
2228: \Asp(\Xhat_P;\AA\inv(\mathscr N_P(\RR)';\EE))
2229: \end{equation*}
2230: and hence there is a natural morphism, denoted $\o\mapsto \o|_{\Xhat_P}$,
2231: \begin{equation}
2232: \Asp(\Xhat;\EE)\longrightarrow
2233: \ihat_{P*}\Asp(\Xhat_P;\AA\inv(\mathscr N_P(\RR)';\EE))\ .
2234: \label{eqnRestrictclosedSpecialForms}
2235: \end{equation}
2236: 
2237: \subsection{}
2238: \label{ssectResolutionBySpecialForms}
2239: A theorem of Nomizu \cite{refnNomizu} and van Est \cite{refnvanEst} implies
2240: that the natural inclusion of complexes of differential forms induces an
2241: isomorphism
2242: \begin{equation*}
2243: H(A\inv(\mathscr N_P(\RR)';\EE))\cong H(\mathscr N_P(\RR)';\EE);
2244: \end{equation*}
2245: from this, the Poincar\'e lemma, and \eqref{eqnRestrictSpecialForms} one
2246: sees that
2247: 
2248: \begin{lem*}
2249: \textup(\cite[\S13]{refnGoreskyHarderMacPherson}\textup) There is a
2250: natural isomorphism $\Asp(\Xhat;\EE)\tildearrow \i_{G*}\EE$ in
2251: $\Derived_{\X}(\Xhat)$ for $E\in\R(G)$.
2252: \end{lem*}
2253: 
2254: \subsection{}
2255: Recall that $C(\n_P;E)$ denotes the complex $\Hom(\bigwedge\n_P;E)$ with
2256: differential as in \cite[I, \S1.1]{refnBorelWallach}.  Its cohomology
2257: $H(\n_P;E)$ is a graded $L_P$-module; let $\HH(\n_P;E)$ denote the
2258: corresponding graded locally constant sheaf on $X_P$.
2259: 
2260: \begin{lem}
2261: \label{ssectNilmanifoldCohomologyBundle}
2262: \textup{(compare
2263: \cite[\S12]{refnGoreskyHarderMacPherson})} There is a
2264: natural quasi-isomorphism on $X_P$\textup:
2265: \begin{equation*}
2266: h_P\colon \AA\inv(\mathscr N_P(\RR)';\EE) \longrightarrow \HH(\n_P;E).
2267: \end{equation*}
2268: \end{lem}
2269: \begin{proof}
2270: Choose a basepoint $x\in D$ and let $L_{P,x}\subseteq P$ denote the lift of
2271: $L_P$ stable under the Cartan
2272: involution associated to $x$.  $L_P$ acts on $C(\n_P;E)$ through
2273: the coadjoint action of $L_{P,x}$; we
2274: denote by $C(\n_P;E)_x$ the resulting complex of $L_P$-modules and by
2275: $\CC(\n_P;E)_x$ the corresponding complex
2276: of locally constant sheaves.  
2277: 
2278: The quasi-isomorphism $h_P$ is given by a composition of quasi-isomorphisms
2279: \begin{equation}
2280: \AA\inv(\mathscr N_P(\RR)';\EE) \overset{h_x}{\longrightarrow}
2281: \CC(\n_P;E)_{x} \longrightarrow
2282: \HH(\n_P;E)\ .\label{eqnToCohomology}
2283: \end{equation}
2284: The first map is induced by an isomorphism
2285: \cite[(12.13)]{refnGoreskyHarderMacPherson} of the underlying
2286: $\G_{L_P}$-modules.  Namely, transfer a form from $\mathscr N_P(\RR)'$ to
2287: $N_P(\RR)'$ via the isomorphism $n_x'$ and evaluate the form at the
2288: identity; one may use \S\ref{ssectActionOnNilmanifoldGivenBasepoint} to
2289: verify this respects the $\G_{L_P}$-action.  For the second map of
2290: \eqref{eqnToCohomology}, note that if $E$ were irreducible, a theorem of
2291: Kostant \cite{refnKostant} says that each irreducible component of
2292: $H(\n_P;E)$ occurs in both $H(\n_P;E)$ and $C(\n_P;E)_x$ with multiplicity
2293: one.  Thus for general regular $E$ there is a unique map $C(\n_P;E)_x\to
2294: H(\n_P;E)$ in $\Complex(L_P)$ inducing the identity on cohomology.  This
2295: induces the second map in \eqref{eqnToCohomology}.
2296: 
2297: If $x'$ is an another basepoint then $L_{P,x'}=v L_{P,x}v^{-1}$ for some
2298: $v\in N_P(\RR)$.  Thus there is an isomorphism $\Ad^*(v)\colon  C(\n_P;E)_{x}
2299: \tildearrow C(\n_P;E)_{x'}$ in $\Complex(L_P)$.  One may check from the
2300: definition of $h_x$ that the left triangle of
2301: \begin{equation*}
2302: \xymatrix{
2303: {} & {\CC(\n_P;E)_{x}} \ar[dd]_{\Ad^*(v)} \ar[rd] & {} \\
2304: {\AA\inv(\mathscr N_P(\RR)';\EE)} \ar[ru]^{h_{x}} \ar[rd]_{h_{x'}} & {} &
2305: {\HH(\n_P;E)} \\
2306: {} & {\CC(\n_P;E)_{x'}} \ar[ru] & {}
2307: }
2308: \end{equation*}
2309: commutes; since $\Ad^*(v)$ induces the identity on cohomology, the right
2310: triangle commutes by the uniqueness noted above.  Thus $h_P$ is independent of
2311: the basepoint $x$.
2312: 
2313: The verification of naturality for each map in \eqref{eqnToCohomology} is
2314: left to the reader.
2315: \end{proof}
2316: 
2317: \subsection{}
2318: \label{ssectRestrictSpecialForms}
2319: This lemma together with equation \eqref{eqnRestrictclosedSpecialForms}
2320: yields the
2321: 
2322: \begin{cor*}
2323: There is a natural morphism
2324: \begin{equation*}
2325: k_P\colon \Asp(\Xhat;\EE) \longrightarrow \ihat_{P*}\Asp(\Xhat_P;\HH(\n_P;E))
2326: \end{equation*}
2327: given by $\o\mapsto \Asp(\Xhat_P;h_P)(\o|_{\Xhat_P})$ such that
2328: $\ihat_P^*(k_P)$ is a quasi-isomorphism.
2329: \end{cor*}
2330: 
2331: \subsection{Proof of Theorem
2332: \ref{ssectRealizationLModules}\itemref{itemExistenceOfBasicRealizationFunctors}}
2333: Apply Lemma~\ref{ssectResolutionBySpecialForms},
2334: Lemma~\ref{ssectNilmanifoldCohomologyBundle}, and Corollary
2335: ~\ref{ssectRestrictSpecialForms} with $X$ and $G$ replaced by $X_Q$ and
2336: $L_Q$ respectively.  We obtain for all $P\le Q$ and all $E_Q\in\R(L_Q)$
2337: natural quasi-isomorphisms
2338: \begin{gather*}
2339: \Asp(\Xhat_Q;\EE_Q)\longrightarrow \i_{Q*}\EE_Q\ , \\
2340: h_{PQ}\colon \AA\inv(\mathscr N_P^Q(\RR)';\EE_Q) \longrightarrow
2341: \HH(\n_P^Q;E_Q)\ ,
2342: \end{gather*}
2343: and a natural morphism
2344: \begin{equation*}
2345: k_{PQ}\colon \ihat_{Q*}\Asp(\Xhat_Q;\EE_Q) \longrightarrow
2346: \ihat_{P*}\Asp(\Xhat_P;\HH(\n_P^Q;E_Q))\ ,
2347: \end{equation*}
2348: where $k_{PQ}(\o) = \Asp(\Xhat_P;h_{PQ})(\o|_{\Xhat_P})$ and
2349: $\ihat_P^*(k_{PQ})$ is a quasi-isomorphism.  The proposition follows if we
2350: set $\A_P(E_P) = \Asp(\Xhat_P;\EE_P)$.  \qed
2351: 
2352: \subsection{}
2353: \label{ssectOtherRealizationsLsheaves}
2354: There are other choices of $\{\A_P\}_{P\in\Pl}$ that yield different
2355: incarnations of the same realization $\Sheaf_W$.  Consider for simplicity
2356: just $P=G$.  Then to avoid the boundary conditions of special differential
2357: forms at the expense of complicating the combinatorics, one may set
2358: $\A_G(E)$ equal to the resolution of $\i_{G*}\EE$ by {\itshape
2359: combinatorial forms\/}
2360: \begin{equation*}
2361: \Acomb(\Xhat;\EE) = \bigoplus_{P\le Q}
2362: \A((\G_P/\G_{N_Q})\back(\overline{N_Q(\RR)\back
2363: e_P});\AA\inv(\mathscr N_Q(\RR)';\EE)).
2364: \end{equation*}
2365: As another example, if $E$ has a $\QQ$-structure one may set $\A_G(E)$
2366: equal to the resolution $\i_{G*}\mathbf I(\EE)$ defined in
2367: \cite[(27.9)]{refnGoreskyHarderMacPherson}; in this way one obtains a
2368: incarnation $\A_W$ defined over $\QQ$ for $\L$-modules defined over $\QQ$.
2369: We will not go into further details here.
2370: 
2371: \section{Example: the Intersection Cohomology $\L$-module}
2372: \label{sectIntersectionCohomologyLmodule}
2373: Let $W\subseteq \Xhat$ be an admissible subspace.
2374: 
2375: \subsection{}
2376: Recall that there are truncation functors on $\Complex(L_P)$ given by
2377: \begin{align*}
2378: \t^{\leqslant n}C\colon &\quad \dots\to C^{n-1} \to \ker d_C^n \to 0\to
2379: \cdots\ ,\\
2380: \t^{> n}C\colon &\quad \dots\to 0 \to C^{n+1}/\Im d_C^n \to C^{n+2}\to
2381: \cdots\ .
2382: \end{align*}
2383: There is a natural short exact sequence $0\to \t^{\leqslant n}C \to C \to
2384: \t^{> n}C \to 0$ whose morphisms induce isomorphisms
2385: \begin{equation*}
2386: H^i(\t^{\leqslant n}C) = \begin{cases} H^i(C) & i\le n, \\
2387:                                 0      & i>n,
2388:                          \end{cases}
2389: \qquad\text{and}\qquad
2390: H^i(\t^{> n}C) = \begin{cases} 0       & i\le n, \\
2391:                         H^i(C)  & i>n.
2392:                   \end{cases}
2393: \end{equation*}
2394: The exact sequence also implies there are natural quasi-isomorphisms
2395: \cite[1.7.5]{refnKashiwaraSchapira}
2396: \begin{equation}
2397: M(\t^{\leqslant n}C\!\to\! C) \tildearrow \t^{> n}C \qquad \text{and}
2398: \qquad \t^{\leqslant n} C \tildearrow M(C\!\to\! \t^{> n}C)[-1]
2399: \label{eqnTruncationIsCone}
2400: \end{equation}
2401: which suggests how to extend these functors to $\R(\L_W)$.
2402: 
2403: \subsection{}
2404: \label{ssectLocalDegreeTruncationLmodule}
2405: For $Q\in\Pl(W)$ consider the composition $\M\to \i_{Q*}\i_Q^*\M \to
2406: \i_{Q*}\t^{> n}\i_Q^*\M$ of natural maps (the first being the adjunction
2407: morphism implied by \eqref{eqnAdjointRelations}).  Define a functor
2408: $\t_Q^{\leqslant n}$ on $\R(\L_W)$ by
2409: \begin{equation*}
2410: \t_Q^{\leqslant n}\M = M(\M\to \i_{Q*}\t^{> n}\i_Q^*\M)[-1]\ ;
2411: \end{equation*}
2412: \S\ref{ssectMappingCone} shows there is a natural morphism $\t_Q^{\leqslant
2413: n}\M\to \M$.  By \eqref{eqnTruncationIsCone} there is a natural quasi-isomorphism
2414: \begin{equation}
2415: \i_P^*\circ \t_Q^{\leqslant n} \cong 
2416:                 \begin{cases} \t^{\leqslant n} \circ \i_Q^* & \text{if $P=Q$,} \\
2417:                        \i_P^*                        & \text{if $P\nleq Q$.}
2418:                 \end{cases} \label{eqnLocalCohomLocalTruncation}
2419: \end{equation}
2420: If $P<Q$ a simple formula like \eqref{eqnLocalCohomLocalTruncation} does not
2421: hold.  Instead there is a short exact sequence
2422: \begin{equation*}
2423: 0 \to H(\n_P^Q;\t^{> n}\i_Q^*\M[-1]) \to \i_P^* \t_Q^{\leqslant n} \M
2424: \to \i_P^*\M \to 0 \ ;
2425: \end{equation*}
2426: since $H(\n_P^Q;\cdot)$ is degree nondecreasing we at least have a
2427: quasi-isomorphism
2428: \begin{equation}
2429: \t^{\leqslant n} \circ \i_P^* \circ \t_Q^{\leqslant n} \cong \t^{\leqslant n}
2430: \circ \i_P^*\ .
2431: \label{eqnTruncatedLocalCohomLocalTruncation}
2432: \end{equation}
2433: It is also clear that
2434: \begin{equation}
2435: \t_P^{\leqslant n}\circ \t_Q^{\leqslant n}=\t_Q^{\leqslant n}\circ
2436: \t_P^{\leqslant n} \quad \text{if $P\nless Q$ and $Q\nless P$.}
2437: \label{eqnCommuteLocalTruncation}
2438: \end{equation}
2439: 
2440: \subsection{Degree truncation of $\L$-modules}
2441: \label{ssectDegreeTruncationLmodules}
2442: The functor $\t^{\leqslant n}$ on $\R(\L_W)$ is defined as
2443: \begin{equation*}
2444: \t^{\leqslant n} = \t_{Q_1}^{\leqslant n}\circ \t_{Q_2}^{\leqslant
2445: n} \circ \dots \circ\t_{Q_N}^{\leqslant n}\ ,
2446: \end{equation*}
2447: where we write $\Pl(W)=\{Q_1,\dots, Q_N\}$ so that $Q_i\le Q_j$ implies
2448: $i\le j$; the functor is independent of the choice of ordering on
2449: $\Pl(W)$ by \eqref{eqnCommuteLocalTruncation}.  There is a natural morphism
2450: $\t^{\leqslant n}\M\to \M$.  From
2451: \eqref{eqnLocalCohomLocalTruncation}--\eqref{eqnCommuteLocalTruncation} it
2452: follows that there is a quasi-isomorphism
2453: \begin{equation*}
2454: \i_Q^*\circ \t^{\leqslant n} \cong \t^{\leqslant n} \circ \i_Q^*\ .
2455: \end{equation*}
2456: 
2457: By applying $\i_Q^*$ and using \eqref{eqnTruncationIsCone} it is easy to
2458: verify the
2459: \begin{lem*}
2460: Let $W$ be an admissible space and let $\Sheaf_W\colon \R(\L_W)\to
2461: \Derived_{\X}(W)$ be a realization functor from Theorem
2462: ~\textup{\ref{ssectRealizationLModules}}.  Then $\Sheaf_W\circ
2463: \t^{\leqslant n} \cong \t^{\leqslant n} \circ\Sheaf_W$.
2464: \end{lem*}
2465: 
2466: \begin{rem*}
2467: The isomorphism does not hold on the level of complexes.  The point is
2468: that $\t^{\leqslant n}$ on $\L$-modules is defined ``externally'', via a
2469: mapping cone, whereas  $\t^{\leqslant n}$ on complexes of sheaves is defined
2470: ``internally'', as a sub-object.
2471: \end{rem*}
2472: 
2473: \subsection{Intersection cohomology $\L$-module}
2474: \label{ssectIntersectionCohomology}
2475: Let $E$ be a regular $G$-module and let $p$ be an ordinary perversity, that
2476: is, a function $p\colon \{2,\dots,\dim X\}\to \ZZ$ such that $p(2)=0$, and
2477: $p(k+1) = p(k)$ or $p(k)+1$.  The {\itshape intersection cohomology
2478: $\L_{\Xhat}$-module $\IpC(E)$ \textup(with perversity $p$ and coefficients
2479: $E$\/\textup)} is defined as follows.  Set
2480: \begin{equation*}
2481: \M_X = E
2482: \end{equation*}
2483: For an admissible space $W\subseteq \Xhat$ strictly containing $X$ let
2484: $P\in \Pl(W)$ be minimal both with respect to the partial ordering on
2485: $\Pl(W)$ and to $\dim X_P$.  Let $\j_P\colon W\setminus X_P\hookrightarrow W$
2486: denote the inclusion.  Define the $\L_W$-module $\M_W$ inductively by
2487: setting
2488: \begin{equation}
2489: \M_W = \t^{\leqslant p(\codim_{\Xhat}X_P)}\j_{P*} \M_{W\setminus
2490: X_P}\ . \label{eqnIntersectionCohomologyLmodule}
2491: \end{equation}
2492: Finally set $\IpC(E)=\M_{\Xhat}$; there is a natural morphism $\IpC(E) \to
2493: \i_{G*}E$.  Note that since $p$ is nondecreasing one may replace
2494: $\t^{\leqslant p(\codim_{\Xhat}X_P)}$ in
2495: \eqref{eqnIntersectionCohomologyLmodule} by $\t_P^{\leqslant
2496: p(\codim_{\Xhat}X_P)}$.
2497: 
2498: Let $\EE= D\times_\G E$ be the locally constant sheaf on $X$ induced by $E$
2499: and let $\IpC(\Xhat;\EE)$ be the corresponding intersection cohomology sheaf
2500: \cite{refnGoreskyMacPhersonIHTwo} in $\Derived_{\X}(\Xhat)$.  The following
2501: proposition is immediate from the definitions and the fact that
2502: $\Sheaf_{\Xhat}$ commutes with the standard functors and with truncation
2503: (Theorem
2504: ~\ref{ssectRealizationLModules}\itemref{itemExistenceOfRealizationFunctors}
2505: and Lemma ~\ref{ssectDegreeTruncationLmodules}).
2506: 
2507: \begin{prop*}
2508: There is a natural isomorphism $\Sheaf_{\Xhat}(\IpC(E)) \cong
2509: \IpC(\Xhat;\EE)$.
2510: \end{prop*}
2511: 
2512: \subsection{Local intersection cohomology}
2513: \label{ssectLocalIntersectionCohomology}
2514: We wish to give a formula for the local cohomology at $P$ of $\IpC(E)$.
2515: There are two ingredients to the formula.  The first is the nilpotent
2516: cohomology $H(\n_P;E)$ and its decomposition $ \bigoplus_{w\in W_P}
2517: H(\n_P;E)_w $ from Kostant's theorem as in \S\ref{ssectKostantsTheorem}.
2518: The second ingredient is a certain combinatorial invariant
2519: $I_{p_w}H(c(|\D_P|))$, which we now describe.
2520: 
2521: Let $|\D_P|$ denote the geometric simplex with vertices indexed by the
2522: elements of $\D_P$.  It has a natural stratification given by the
2523: decomposition into open faces.  The strata correspond to parabolic
2524: $\QQ$-subgroups $Q\gneq P$ (that is, to nonempty subsets $\D_P^Q\subseteq
2525: \D_P$) and are denoted $|\D_P^Q|^\circ$.  We give the cone $c(|\D_P|)$ the
2526: induced stratification---the strata consist of the open cones on
2527: $|\D_P^Q|^\circ$ for $Q\gneq P$ together with the cone point (indexed by
2528: $P$).  Any constructible subset (that is, any union of open faces) of
2529: $|\D_P|$ or $c(|\D_P|)$ is given the induced stratification.
2530: 
2531: Given an ordinary perversity $p$ and an element $w\in W$, define  $p_w\colon 
2532: \Pl\setminus \{G\} \to \ZZ$ by
2533: \begin{equation*}
2534: p_w(Q) = p(\dim \n_Q + \#\D_Q) - \l(w_Q),
2535: \end{equation*}
2536: where $w_Q$ denotes the projection of $w$ to the second factor of $W=W^Q
2537: W_Q$.  This defines an integer-valued function on the singular strata of
2538: $c(|\D_P|)$ or $|\D_P|$ (a perversity in the sense of
2539: \cite{refnBeilinsonBernsteinDeligne}).  Let $I_{p_w}H(U)$ denote the
2540: corresponding intersection cohomology group over $\ZZ$ for any open
2541: constructible subset $U$ of $|\D_P|$ or $c(|\D_P|)$.  (Unlike the
2542: intersection cohomology for an ordinary perversity $p$, the group here
2543: depends on a choice of stratification; we always use the stratification
2544: fixed above.)  Note that if $P\le Q$, then $w_Q = (w_P)_Q$ and so $p_w$ and
2545: $p_{w_P}$ agree on the strata of $|\D_P|$.
2546: 
2547: \begin{prop*}
2548: $H(\i_P^*\IpC(E)) \cong \bigoplus_{w\in W_P} H(\n_P;E)_w\otimes
2549: I_{p_w}H(c(|\D_P|))$ where $L_P$ acts trivially on the second factor.
2550: \end{prop*}
2551: \begin{proof}
2552: There is a Mayer-Vietoris spectral sequence that abuts to the link
2553: cohomology $H(\i_P^*\j_{P*}\j_P^*\IpC(E))$ (Lemma ~
2554: \ref{ssectRelativeLocalCohomologySupportsSS} with $Q'=G$ and $Q=P$) with
2555: \begin{equation*}
2556: E_1^{p,\cdot}= \bigoplus_{\#\D_P^{\tilde P}=p+1} H( \n_P^{\tilde P}; H(\i_{\tilde P}^*\IpC(E)))\ .
2557: \end{equation*}
2558: By induction we may assume the proposition is true for ${\tilde P}>P$ and thus we
2559: can compute
2560: \begin{equation}
2561: \begin{split}
2562: E_1^{p,\cdot} &\cong \bigoplus_{\#\D_P^{\tilde P}=p+1}
2563:           H( \n_P^{\tilde P};\bigoplus_{w_{\tilde P}\in W_{\tilde P}} H(\n_{\tilde P};E)_{w_{\tilde P}}\otimes 
2564:           I_{p_{w_{\tilde P}}}H( c(|\D_{\tilde P}|) ) )\\
2565:           &\cong \bigoplus_{\#\D_P^{\tilde P}=p+1} \bigoplus_{w_{\tilde P}\in 
2566:           W_{\tilde P}} \bigoplus_{w^{\tilde P}_P\in W^{\tilde P}_P}
2567:           H(\n_P^{\tilde P}; H(\n_{\tilde P};E)_{w_{\tilde P}})_{w^{\tilde P}_P} \otimes 
2568:           I_{p_{w_{\tilde P}}}H( c(|\D_{\tilde P}|) ) \\
2569:           &\cong \bigoplus_{w\in W_P} H(\n_P;E)_w \otimes
2570:           \bigoplus_{\#\D_P^{\tilde P}=p+1} I_{p_w}H( c(|\D_{\tilde P}|) )\ .
2571: \end{split}  \label{eqno}
2572: \end{equation}
2573: 
2574: On the other hand, there is an analogous Mayer-Vietoris spectral sequence
2575: abutting to $I_{p_w}H(|\D_P|)$.  Namely cover $|\D_P|$ by the open stars
2576: $U_\al$ of the vertices $\al\in\D_P$.  For a parabolic $\QQ$-subgroup
2577: ${\tilde P}> P$, the intersection $U_{\tilde P}=\bigcap_{\al\in
2578: \D_P^{\tilde P}} U_\al$ is the open star of the open face $|\D_P^{\tilde
2579: P}|^\circ$.  For this spectral sequence
2580: \begin{equation}
2581: \widetilde E_1^{p,\cdot} =\bigoplus_{ \#\D_P^{\tilde P}=p+1} I_{p_w}H(U_{\tilde P}) \cong 
2582: \bigoplus_{ \#\D_P^{\tilde P}=p+1} I_{p_w}H(c(|\D_{\tilde P}|))\ . \label{eqnp}
2583: \end{equation}
2584: Comparing \eqref{eqno} ~and \eqref{eqnp} we see that $E_1^{p,\cdot} \cong
2585: \bigoplus_{w\in W_P} H(\n_P;E)_w \otimes \widetilde E_1^{p,\cdot}$.  This
2586: isomorphism induces isomorphisms at all stages of the spectral sequences
2587: and we find that
2588: \begin{equation*}
2589: H(\i_P^*\j_{P*}\j_P^*\IpC(E))\cong \bigoplus_{w\in W_P} 
2590: H(\n_P;E)_w \otimes I_{p_w}H(|\D_P|)\ .
2591: \end{equation*}
2592: In order to obtain $H(\i_P^*\IpC(E))$, this link cohomology is to be
2593: truncated in degrees greater than $p(\codim_{\Xhat} X_P) = \l(w) + p_w(P)$,
2594: which corresponds exactly to replacing the second factor by
2595: $I_{p_w}H(c(|\D_P|))$.
2596: \end{proof}
2597: 
2598: \section{Example: the Weighted Cohomology $\L$-module}
2599: \label{sectWeightCohomologyLmodule}
2600: 
2601: \subsection{}
2602: The reference for this subsection is \cite[
2603: ~\S9]{refnGoreskyHarderMacPherson} though we give a slightly different
2604: description.
2605: 
2606: Let $P\in \Pl$ and let $X(S_P^G)$ denote the lattice of rational characters
2607: on $S_P^G$.  For $\al\in\D_P$, let $R_\al\ge P$ be the maximal parabolic
2608: $\QQ$-subgroup having type $\D_P\setminus \{\al\}$ with respect to $P$.  We
2609: may canonically identify $X(S_{R_\al}^G)\cong \ZZ$ so that the unique
2610: element of $\D_{R_\al}$ corresponds to a positive integer.  Consider the
2611: map $X(S_P^G)\hookrightarrow \prod_{ \al\in\D_P} X(S_{R_\al}^G) \cong
2612: \ZZ^{\D_P}$ defined by $\chi \mapsto (\chi|_{S_{R_\al}^G})_{\al\in\D_P}$.
2613: The image is a sublattice of finite index; after tensoring with $\RR$ we
2614: have an isomorphism $\phi_P\colon \sa_P^{G*} = X(S_P^G)\otimes_\ZZ \RR
2615: \tildearrow \RR^{\D_P}$.  For $P=P_0$ minimal we omit the subscript.
2616: 
2617: Define a partial ordering on $\sa_P^{G*}$ by declaring $\zeta\ge \eta$ if
2618: and only if $\phi_P(\zeta)\ge \phi_P(\eta)$ component-wise.  Thus if
2619: $\lsp+\sa_P^{G*}$ denotes the real convex cone generated by all
2620: $\al\in\D_P$, we have
2621: \begin{equation}
2622: \begin{aligned}
2623: \zeta\ge \eta &\qquad\Longleftrightarrow\qquad \zeta-\eta\in
2624: 	\lsp+\sa_P^{G*} \\
2625:  &\qquad\Longleftrightarrow\qquad \langle\zeta,\b\spcheck\rangle\ge
2626: 	\langle\eta,\b\spcheck\rangle \qquad \text{for all $\b \in \Dhat_P$.}
2627: \end{aligned}
2628: \label{eqnWeightOrdering}
2629: \end{equation}
2630: 
2631: If $P\le Q$ and $\zeta \in \sa_P^{G*}$, let $\zeta_Q = \zeta|_{\sa_Q^G}$
2632: denote the restriction.  We have a commutative diagram
2633: \begin{equation*}
2634: \begin{CD}
2635: \sa_P^{G*}  @>\phi_P>> \RR^{\D_P} \\
2636: @VVV     @VVV \\
2637: \sa_Q^{G*} @>\phi_Q>> \RR^{\D_Q}
2638: \end{CD}
2639: \end{equation*}
2640: where the left-hand vertical map is restriction and the right-hand vertical
2641: map is the projection $\RR^{\D_P} = \RR^{\D_P^Q}\times \RR^{\D_P\setminus
2642: \D_P^Q}\to \RR^{\D_P\setminus \D_P^Q} \cong \RR^{\D_Q}$.
2643: 
2644: For a regular $L_P$-module $E$ we write
2645: \begin{equation*}
2646: E= \bigoplus_{\chi\in X(S_P^G)} E_\chi
2647: \end{equation*}
2648: where $E_\chi$ is the submodule on which $S_P^G$ acts via $\chi$.  Given
2649: $\eta \in \sa^{G*}$, there are {\itshape weight truncation functors\/} on
2650: $\R(L_P)$ defined by
2651: \begin{equation*}
2652: \t^{\geqslant \eta}E = \bigoplus_{\chi \ge \eta_P} E_\chi  \quad\text{and}\quad
2653: \t^{\ngeqslant \eta}E = \bigoplus_{\chi \ngeq \eta_P} E_\chi \ ;
2654: \end{equation*}
2655: the element $\eta$ will be called a \emph{weight profile}.  (In
2656: \cite{refnGoreskyHarderMacPherson} a weight profile is required to satisfy
2657: $\phi(\eta)\in (\ZZ+\frac12)^{\D}$; this ensures that any truncation
2658: functor $\t^{\geqslant \eta}$ arises from a unique weight profile $\eta$.
2659: We will not assume this, but we will feel to replace $\eta$ by $\eta'$
2660: provided $\t^{\geqslant \eta}$ is unchanged.)  Clearly we have a split
2661: short exact sequence $0\to \t^{\geqslant \eta}E \to E \to \t^{\ngeqslant
2662: \eta}E \to 0$.
2663: 
2664: \subsection{Weight truncation of $\L$-modules}
2665: \label{ssectWeightTruncationLmodule}
2666: Consider $P\le Q\in \Pl$ and let $E$ be a regular $L_Q$-module.  The
2667: subtorus $S_Q^G\subseteq S_P^G$ acts on $H(\n_P^Q; E_\chi)$ via the
2668: character $\chi$.  Thus
2669: \begin{equation}
2670: \t^{\geqslant \eta} H(\n_P^Q; E) \subseteq H(\n_P^Q; \t^{\geqslant \eta}
2671: E)\ . \label{eqnWeightNonIncreasing}
2672: \end{equation}
2673: 
2674: Given a weight profile $\eta$ we define local weight truncation functors
2675: $\t_Q^{\geqslant \eta}$ on $\R(\L_W)$ via a mapping cone
2676: \begin{equation}
2677: \t_Q^{\geqslant \eta}\M = M(\M\to \i_{Q*}\t^{\ngeqslant
2678: \eta}\i_Q^*\M)[-1]\ ;
2679: \end{equation}
2680: these are analogous to the local degree truncation functors in
2681: \S\ref{ssectLocalDegreeTruncationLmodule}.  The only change is in proving
2682: the analogue of \eqref{eqnTruncatedLocalCohomLocalTruncation}.  Instead of
2683: the fact that $H(\n_P^Q;\cdot)$ is degree nondecreasing, we use that it is
2684: weight nonincreasing by \eqref{eqnWeightNonIncreasing}; we also use that
2685: $\t^{\geqslant \eta}$ is an exact functor.
2686: 
2687: The weight truncation functor $\t^{\geqslant \eta}$ on $\R(\L_W)$ is
2688: defined as a composition of these local truncation functors as in
2689: \S\ref{ssectDegreeTruncationLmodules}.  There is a natural morphism
2690: $\t^{\geqslant \eta}\M \to \M$ and a quasi-isomorphism
2691: \begin{equation}
2692: \i_Q^*\circ \t^{\geqslant \eta} \cong  \t^{\geqslant \eta}\circ \i_Q^* \ .
2693: \label{eqnWeightTruncationPullbackToStratumCommute}
2694: \end{equation}
2695: 
2696: \subsection{Weighted cohomology $\L$-module}
2697: Let $E$ be a regular $G$-module and let $\eta$ be a weight profile.  The
2698: {\itshape weighted cohomology $\L_{\Xhat}$-module $\WnC(E)$ \textup(with
2699: weight profile $\eta$ and coefficients $E$\/\textup)} is defined as
2700: \begin{equation*}
2701: \WnC(E) = \t^{\geqslant \eta} \i_{G*}E \ .
2702: \end{equation*}
2703: 
2704: There is a natural morphism $\WnC(E) \to \i_{G*}E$.  Now let
2705: $\WnC(\Xhat;\EE)$ be the corresponding weighted cohomology sheaf
2706: \cite{refnGoreskyHarderMacPherson} in $\Derived_{\X}(\Xhat)$.
2707: 
2708: \begin{prop*}
2709: There is a natural isomorphism $\Sheaf_{\Xhat}(\WnC(E)) \cong \WnC(\Xhat;\EE)$.
2710: \end{prop*}
2711: 
2712: \subsection{Local weighted cohomology}
2713: \label{ssectLocalWeightedCohomology}
2714: The formula for the local cohomology at $P$ of $\WnC(E)$ is considerably
2715: simpler than that for $\IpC(E)$ since the former is defined by a simple
2716: truncation operation as opposed to the inductive truncation procedure of
2717: the latter.  The following proposition is an immediate consequence of
2718: \eqref{eqnWeightTruncationPullbackToStratumCommute} and the definitions.
2719: 
2720: \begin{prop*}
2721: $H(\i_P^*\WnC(E)) \cong \t^{\geqslant \eta}
2722: H(\i_P^*\i_{G*}E) \cong \t^{\geqslant \eta} H(\n_P;E)$.
2723: \end{prop*}
2724: 
2725: \begin{rem}
2726: In order to rephrase this in a form analogous to Proposition
2727: ~\ref{ssectLocalIntersectionCohomology}, let $E$ be irreducible and have
2728: highest weight $\lambda$ with respect to a Cartan subalgebra of $\levi_P$
2729: and a positive system $\Phi^+(\mathfrak g_\CC,\h_\CC)$ containing
2730: $\Phi(\n_{P\CC},h_{\CC})$.  Then by Kostant's theorem
2731: \S\ref{ssectKostantsTheorem} we have
2732: \begin{equation}
2733: H(\i_P^*\WnC(E)) \cong \bigoplus_{w\in W_P} H(\n_P;E)_w\otimes
2734: \t^{\geqslant \eta_P -(w(\lambda +\hsr)-\hsr)_P}\CC
2735: \end{equation}
2736: the second factor is either $0$ or $\CC$ with the trivial action of $L_P$.
2737: As Rapoport pointed out to me, the second factor here depends on both $w$
2738: and $E$, unlike $I_{p_w}H(c(|\D_P|))$ of Proposition
2739: ~\ref{ssectLocalIntersectionCohomology} which is independent of $E$.
2740: \end{rem}
2741: 
2742: \section{Micro-support of an $\L$-module}
2743: \label{sectMicroSupport}
2744: In this section we define the micro-support of an $\L$-module.  This is an
2745: analogue of the notion for sheaves \cite{refnKashiwaraSchapira}; the
2746: analogy is not precise, partially because it is most natural here to
2747: restrict certain modules to be isomorphic to the complex conjugate of their
2748: contragredient.
2749: 
2750: \subsection{Notation}
2751: \label{ssectMicroSupportNotation}
2752: Let $W$ be an admissible space possessing a unique maximal stratum $X_R$ and
2753: set
2754: \begin{equation*}
2755: \IrrRep(\L_W)= \coprod_{P\in \Pl(W)} \IrrRep(L_P)\ .
2756: \end{equation*}
2757: For an $L_P$-module $V\in\IrrRep(\L_W)$, define parabolic $\QQ$-subgroups
2758: $Q^{\prime W}_V\ge Q^W_V\ge P$ by
2759: \begin{align*}
2760: \D_{P}^{Q^W_V} &= \{\,\al\in \D_{P}^R \mid
2761: \langle\xi_V+\hsr,\al\spcheck\rangle <0\,\}\ ,\\
2762: \D_{P}^{Q^{\prime W}_V} &= \{\,\al\in \D_{P}^R \mid
2763: \langle\xi_V+\hsr,\al\spcheck\rangle \le 0\,\}\ .
2764: \end{align*}
2765: Note that $Q_V^W$ and $Q^{\prime W}_V$ depend on $W$ through the parabolic
2766: $\QQ$-subgroup $R$ indexing its maximal stratum.  When $W=\Xhat$ we
2767: simply write $Q_V$ and $Q'_V$.
2768: 
2769: \begin{defn}
2770: \label{ssectMicroSupport}
2771: The {\itshape weak micro-support\/} $\mS_w(\M)$ of an $\L_W$-module $\M$ is
2772: the set of all $V\in \IrrRep(\L_W)$ such that
2773: \begin{enumerate}
2774: \item $H(\i_{P}^* \ihat_{Q}^! \M)_V \neq 0$ for some
2775: $Q\in [Q_V^W, Q^{\prime W}_V]$.
2776: \label{itemMicroSupportQType}
2777: \setcounter{saveenum}{\value{enumi}}
2778: \end{enumerate}
2779: The {\itshape micro-support\/} $\mS_w(\M)$ consists of those $V\in
2780: \mS_w(\M)$ which satisfy in addition
2781: \begin{enumerate}
2782: \setcounter{enumi}{\value{saveenum}}
2783: \item $(V|_{M_P})^*\cong \overline {V|_{M_P}}$.
2784: \label{itemMicroSupportConjugateSelfcontragedient}
2785: \end{enumerate}
2786: The {\itshape essential micro-support\/} $\emS(\M)$ is the set of
2787: those $V\in\mS(\M)$ such that
2788: \begin{equation}
2789: H(\i_{P}^* \ihat_{{Q_V^W}}^! \M)_V \longrightarrow H(\i_{P}^*
2790: \ihat_{Q^{\prime W}_V}^! \M)_V
2791: \label{eqnMicroNonVanishing}
2792: \end{equation}
2793: is nonzero.
2794: \end{defn}
2795: 
2796: In \S\ref{sectConjugateSelfContragredientFundamentalWeyl} we will discuss
2797: condition \itemref{itemMicroSupportConjugateSelfcontragedient} further;
2798: this will lead in \S\ref{sectAlternateMicroSupport} to alternate forms of
2799: condition \itemref{itemMicroSupportQType} and a clarification of the
2800: relationship between $\mS(\M)$ and $\emS(\M)$.
2801: 
2802: \section{Conjugate Self-contragredient Modules and Fundamental Weyl Elements}
2803: \label{sectConjugateSelfContragredientFundamentalWeyl}
2804: We recall several results due to Borel and Casselman
2805: \cite{refnBorelCasselman}.  The first is an alternate form of Definition ~
2806: \ref{ssectMicroSupport}\itemref{itemMicroSupportConjugateSelfcontragedient}
2807: in terms of highest weights.  The second relates this condition on certain
2808: irreducible components $V$ of $H(\n_P;E)$ to the condition on $E$.
2809: 
2810: \subsection{The involution $\boldsymbol\tau_{\mathbf P}$}
2811: \label{ssectDefineTau}
2812: Let $P$ be a parabolic $\RR$-subgroup and let $V$ be an irreducible regular
2813: $L_P$-module.  We say that $V$ is {\itshape conjugate
2814: self-contragredient\/} if
2815: \begin{equation*}
2816: V^*\cong \overline V
2817: \end{equation*}
2818: as representations of $L_P$.  Clearly this is the case if and only if there
2819: exists a nondegenerate $L_P$-invariant sesquilinear form on $V$, however we
2820: will not use this characterization.  Instead let $\h$ be a Cartan
2821: subalgebra of $\levi_P$ and fix a positive system
2822: $\Phi^+(\levi_{P\CC},\h_\CC)$ of roots.  Borel and Casselman
2823: \cite[\S1]{refnBorelCasselman} construct an involution $\t_P\colon \h^*_\CC\to
2824: \h^*_\CC$ that sends the highest weight of a representation of $L_P$ to its
2825: complex conjugate contragredient.  Thus $V$ is conjugate
2826: self-contragredient if and only if
2827: \begin{equation*}
2828: \t_P(\u) = \u.
2829: \end{equation*}
2830: where $\u$ is the highest weight of $V$.
2831: 
2832: To define $\t_P$, first let $\h^*_\RR$ be the real form of $\h^*_\CC$
2833: spanned by the roots $\Phi(\levi_{P\CC},\h_\CC)$ and the differentials of
2834: rational characters of the center of $L_P$. Let $s^{P}\in W^P$ be the
2835: element of the Weyl group of $\levi_P$ satisfying
2836: $s^{P}(-\Phi^+(\levi_{P\CC},\h_\CC)) =\Phi^+(\levi_{P\CC},\h_\CC)$;
2837: similarly if $c$ denotes complex conjugation of $\h^*_\CC$ with respect to
2838: the original real form $\h^*$ let $s^c\in W^P$ satisfy
2839: $s^c(c\Phi^+(\levi_{P\CC},\h_\CC))=\Phi^+(\levi_{P\CC},\h_\CC)$.  The
2840: automorphism $-s^P$ on $\h^*_\RR$ transforms the highest weight of an
2841: irreducible $L_P$-module to that of its contragredient, while $s^c c$ on
2842: $\h^*_\RR$ transforms the highest weight of an irreducible $L_P$-module to
2843: that of its complex conjugate representation.  Thus the desired
2844: automorphism of $\h^*_\RR$ is
2845: \begin{equation*}
2846: \t_P = (- s^{P}) \circ (s^c c) = (s^c c) \circ (- s^{P})
2847: \end{equation*}
2848: and we extend this $\CC$-linearly to $\h^*_\CC$.  We let $\t_P$ also denote
2849: the dual involution of $\h_\CC$.
2850: 
2851: The involution $\t_P$ depends on the choice of the Cartan subalgebra $\h$
2852: and the positive system of roots.  If $\tilde
2853: \Phi^+(\levi_{P\CC},\h_\CC)=w\Phi^+(\levi_{P\CC},\h_\CC)$ for $w\in W^P$,
2854: then the corresponding involution $\tilde \t_P$ satisfies
2855: \begin{equation}
2856: \tilde \t_P = w \t_P w^{-1} \ .\label{eqnTildeTau}
2857: \end{equation}
2858: 
2859: \subsection{An alternate form of Definition
2860: ~\ref{ssectMicroSupport}\itemref{itemMicroSupportConjugateSelfcontragedient}}
2861: \label{ssectConjugateSelfContragredient}
2862: Now assume that $P$ is defined over $\QQ$.  We may write $\h=\hb_P+\sa_P$
2863: where $\hb_P$ is a Cartan subalgebra of $\m_P$.  Note that $\t_P$ acts by
2864: negation on $\sa_P$ and leaves $\hb_{P\CC}$ invariant.  Let $\hb_{P\CC} =
2865: \hb_{P,1}+\hb_{P,-1}$ be the decomposition into the $+1$ and $-1$
2866: eigenspaces of $\t_P$.
2867: 
2868: The following conditions are equivalent (where $\u$ denotes the highest
2869: weight of $V$):
2870: \begin{gather}
2871: (V|_{M_P})^*\cong \overline{V|_{M_P}}, \text{ that is, $V|_{M_P}$ is
2872: conjugate self-contragredient,} \\
2873: \t_P(\u|_{\hb_P}) = \u|_{\hb_P}, \label{eqnTauInvariant} \\
2874: \u|_{\hb_{P,-1}}=0. \label{eqnEquivalentExpression}
2875: \end{gather}
2876: (In \eqref{eqnTauInvariant} and elsewhere we attempt to lighten the
2877: notation by writing $\u|_{\hb_P}$ for $\u|_{\hb_{P\CC}}$.)
2878: 
2879: For later use we set $\t'_P =-\t_P$.
2880: 
2881: \subsection{Example}
2882: \label{ssectConjugateSelfContragredientExample}
2883: Recall that $\hsr=\hsr^P+\hsr_P$, where $\hsr^P$ is one half the sum of
2884: roots in $\Phi^+(\levi_{P\CC},\h_\CC)$ and $\hsr_P$ is one half the sum of
2885: roots in $\Phi(\n_{P\CC},\h_\CC)$.  The set $\Phi(\n_{P\CC},\h_\CC)$ is
2886: $W^P$-invariant and thus $\hsr_P|_{\hb_P}=0$.  On the other hand,
2887: $\Phi^+(\levi_{P\CC},\h_\CC)$ is $\t_P$-invariant and thus $\t_P(\hsr^P) =
2888: \hsr^P$.  Together these imply
2889: \begin{equation}
2890: \t_P(\hsr|_{\hb_{P}}) = \hsr|_{\hb_{P}} \ .
2891: \label{eqnDeltaConjugateSelfContragredient}
2892: \end{equation}
2893: 
2894: \subsection{Fundamental Cartan subalgebras}
2895: \label{ssectFundamentalCSA}
2896: The familiar situation is when $\t_P=\theta$ for some Cartan involution
2897: $\theta$ of $\levi_P$ and hence $\t'_P$ acts on roots as complex
2898: conjugation.  For a fixed $P$, this can always be arranged by choosing $\h$
2899: and $\Phi^+(\levi_{P\CC},\h_\CC)$ appropriately.  Namely $\t_P=\theta$ if
2900: and only if $\h$ is a $\theta$-stable Cartan subalgebra of $\levi_P$ and
2901: $\Phi^+(\levi_{P\CC},\h_\CC)$ is $\theta$-stable.  A $\theta$-stable
2902: positive system exists if and only if there exists a regular element in the
2903: compact part of $\h$ and hence if and only if the compact part of $\h$ has
2904: maximal dimension.  A Cartan subalgebra with this property for some Cartan
2905: involution is called {\itshape fundamental\/} (or {\itshape maximally
2906: compact\/}) for $\levi_P$; fundamental Cartan subalgebras always exist.
2907: 
2908: In the case that $\t_P=\theta$ the decomposition $\hb_{P\CC} =
2909: \hb_{P,1}+\hb_{P,-1}$ is defined over $\RR$ and corresponds to the Cartan
2910: decomposition into compact and $\RR$-split parts; the condition that
2911: $V|_{M_P}$ is conjugate self-contragredient is equivalent to $\u$
2912: vanishing on the $\RR$-split part of $\hb_P$.
2913: 
2914: \subsection{Equal-rank groups}
2915: \label{ssectEqualRankGroups}
2916: A reductive algebraic group $Z$ defined over $\RR$ will be called {\itshape
2917: equal-rank\/} if $\CCrank Z = \rank K_Z$, where $K_Z$ is a maximal
2918: compact subgroup of $Z(\RR)$.  This holds if and only if a fundamental
2919: Cartan subalgebra is compact and thus if and only if every regular
2920: representation of $Z$ is conjugate self-contragredient.
2921: 
2922: \subsection{Fundamental parabolic subgroups}
2923: \label{ssectFundamentalParabolic}
2924: A Cartan subalgebra $\h$ for $\levi_P$ may of course be identified with a
2925: Cartan subalgebra for $\mathfrak g$ via a lift; we will always assume this
2926: has been done.  Consider the condition on a parabolic $\RR$-subgroup $P$
2927: that a fundamental Cartan subalgebra $\h$ for $\levi_P$ is also fundamental
2928: for $\mathfrak g$.  Such parabolic $\RR$-subgroups exist and a minimal one
2929: is called {\itshape fundamental}.  If $P_0$ is a fundamental parabolic
2930: $\RR$-subgroup, then $L_{P_0}/\lsb\RR S_{P_0}$ is equal-rank; $P_0\neq G$
2931: if and only if $G/\lsb\RR S_G$ is not equal-rank.  Various conditions
2932: equivalent to $P$ containing a fundamental parabolic $\RR$-subgroup are
2933: given in \cite[1.8]{refnBorelCasselman}; here are some others more suitable
2934: for our purposes.
2935: 
2936: \begin{lem*}
2937: Let $P$ be a parabolic $\RR$-subgroup of $G$ and fix a Cartan subalgebra
2938: $\h$ of $\levi_P$.  Let $\Phi^+(\levi_{P\CC},\h_\CC)$ be a positive system
2939: for $\levi_{P\CC}$ and extend it to a positive system $\Phi^+=\Phi^+(\mathfrak
2940: g_\CC,\h_\CC)$ containing $\Phi(\n_{P\CC},\h_\CC)$.  Then the following
2941: conditions are equivalent\textup:
2942: \begin{enumerate}
2943: \item $P$ contains a fundamental parabolic $\RR$-subgroup\textup;
2944: \label{itemFundamental}
2945: \item There exists $w\in W_P$ such that $w\t_G
2946: w^{-1}=\t_P$\textup;
2947: \label{itemTauInvariance}
2948: \item There exists $w\in W_P$ such that $\t_P'$ interchanges $\Phi_w$
2949: and $\Phi(\n_{P\CC},\h_\CC)\setminus \Phi_w$, where $\Phi_w = \{\,\g\in
2950: \Phi^+\mid w^{-1}\g < 0\,\}$\textup;
2951: \label{itemPhiwSwap}
2952: \item There exists $w\in W_P$ such that $w\Phi^+$ is
2953: $\t_P$-stable.
2954: \label{itemwPhiPositiveInvariance}
2955: \end{enumerate}
2956: The conditions in \itemref{itemTauInvariance},
2957: \itemref{itemPhiwSwap} and
2958: \itemref{itemwPhiPositiveInvariance} for a given $w$ are
2959: equivalent.
2960: \end{lem*}
2961: 
2962: \begin{proof}
2963: Let $\t_P$ and $\t_G$ be defined as in \S\ref{ssectDefineTau} with respect
2964: to $\Phi^+(\levi_{P\CC},\h_\CC)$ and $\Phi^+$ respectively.  We choose a
2965: Cartan involution $\theta$ for $\mathfrak g$ which induces one on $\levi_P$
2966: via a $\theta$-stable lift.  We can assume that $\h$ is a fundamental
2967: $\theta$-stable Cartan subalgebra for $\levi_P$ and (by conjugating by an
2968: element of $W^P$) that $\Phi^+(\levi_{P\CC},\h_\CC)$ is $\theta$-stable and
2969: thus that $\t_P=\theta$.
2970: 
2971: \itemref{itemFundamental}$\Rightarrow$\itemref{itemTauInvariance}: Suppose
2972: that $P$ contains a fundamental parabolic $\RR$-subgroup.  Then $\h$ is a
2973: fundamental Cartan subalgebra for $\mathfrak g$ as well so there exists a
2974: $\theta$-stable positive system $\tilde \Phi^+$ for $\mathfrak g$ which
2975: contains $\Phi^+(\levi_{P\CC},\h_\CC)$.  Let $\tilde \t_G$ denote the
2976: operator of \S\ref{ssectDefineTau} with respect to $\tilde\Phi^+$; since
2977: $\tilde\Phi^+$ is $\theta$-stable it satisfies $\tilde \t_G=\theta=\t_P$,
2978: while if $\tilde \Phi^+=w\Phi^+$ for $w\in W$ then $\tilde \t_G = w\t_G
2979: w^{-1}$.  Since both positive systems contain
2980: $\Phi^+(\levi_{P\CC},\h_\CC)$, $w$ must lie in $W_P$.
2981: 
2982: \itemref{itemTauInvariance}$\Rightarrow$\itemref{itemPhiwSwap}: Suppose
2983: $w\t_G w^{-1}=\t_P$ for some $w\in W_P$.  Then $\tilde \t_G=\t_P$ where
2984: $\tilde \t_G$ is with respect to $\tilde \Phi^+=w\Phi^+$.  Decompose
2985: $\Phi(\n_{P\CC},\h_\CC)=\Phi_w \cup (\Phi(\n_{P\CC},\h_\CC)\setminus
2986: \Phi_w)$.  Observe that these subsets correspond to those roots that are
2987: negative (respectively positive) with respect to $\tilde \Phi^+$.  However
2988: $\t_P'$ preserves $\Phi(\n_{P\CC},\h_\CC)$ and since $\t_P'=\tilde \t_G'$
2989: it must interchange the two subsets.
2990: 
2991: \itemref{itemPhiwSwap}$\Rightarrow$\itemref{itemwPhiPositiveInvariance}: If
2992: $w$ is as in \itemref{itemPhiwSwap}, the positive system $w\Phi^+=
2993: \Phi^+(\levi_{P\CC},\h_\CC) \cup -\Phi_w \cup
2994: (\Phi(\n_{P\CC},\h_\CC)\setminus \Phi_w)$ is $\t_P$-stable.
2995: 
2996: \itemref{itemwPhiPositiveInvariance}$\Rightarrow$\itemref{itemFundamental}:
2997: Since $w\Phi^+$ is $\theta$-stable, $\h$ is fundamental for $\mathfrak g$.
2998: \end{proof}
2999: 
3000: \subsection{Fundamental Weyl elements}
3001: \label{ssectFundamentalWeylElement}
3002: An element $w\in W_P$ satisfying $w\t_G w^{-1}=\t_P$ as in Lemma
3003: ~\ref{ssectFundamentalParabolic}\itemref{itemTauInvariance} will be called
3004: a {\itshape fundamental Weyl element\/} (for $P$ in $G$).  More generally,
3005: if $P\le R$ an element $w \in W_P^R$ is fundamental for $P$ in $R$ if
3006: $w\t_R w^{-1}=\t_P$ (equivalently it is fundamental for $P/N_R$ in $L_R$).
3007: Besides the equivalent formulations given in Lemma
3008: ~\ref{ssectFundamentalParabolic}\itemref{itemPhiwSwap} and
3009: \itemref{itemwPhiPositiveInvariance}, here are two basic facts about
3010: fundamental Weyl elements:
3011: \begin{enumerate}
3012: \item A fundamental Weyl element $w\in W_P$ satisfies $\l(w)=\frac12\dim
3013: \n_P$.
3014: \label{itemFundamentalWeylHalf}
3015: \item If $P\le R$ and $w=w^Rw_R\in W^R_P W_R=W_P$, then $w$ is
3016: fundamental for $P$ in $G$ if and only if $w^R$ is fundamental for $P$ in
3017: $R$ and $w_R$ is fundamental for $R$ in $G$.
3018: \label{itemFundamentalWeylHereditary}
3019: \end{enumerate}
3020: Fact \itemref{itemFundamentalWeylHalf} is clear from
3021: Lemma~\ref{ssectFundamentalParabolic}\itemref{itemPhiwSwap}.  For
3022: \itemref{itemFundamentalWeylHereditary}, note that $\t_P$ leaves $\Phi(
3023: \levi_{R\CC},\h_\CC)$ invariant and thus if $w\Phi^+$ is $\t_P$-stable, so
3024: is $w^R\Phi^+( \levi_{R\CC},\h_\CC)= w\Phi^+\cap \Phi(
3025: \levi_{R\CC},\h_\CC)$.  Hence by Lemma
3026: ~\ref{ssectFundamentalParabolic}\itemref{itemwPhiPositiveInvariance},
3027: $w\t_Gw^{-1}=\t_P$ implies $w^R\t_R(w^R)^{-1}=\t_P$ and consequently also
3028: $w_R\t_G w_R^{-1}=\t_R$.  The converse is clear.
3029: 
3030: \begin{lem}
3031: \label{ssectBorelCasselman}
3032: Let $E$ be an irreducible regular $G$-module and let $P$ be a parabolic
3033: subgroup of $G$.  The following conditions are equivalent\textup:
3034: \begin{enumerate}
3035: \item $P$ contains a fundamental parabolic $\RR$-subgroup and
3036: $(E|_{\lsp0G})^*\cong \overline{E|_{\lsp0G}}$\textup;
3037: \label{itemFundamentalGTauInvariant}
3038: \item There exists an irreducible constituent $V$ of $H(\n_P;E)$ satisfying
3039: $(\xi_V+\hsr)|_{\sa^G_P}=0$ and $(V|_{M_P})^*\cong \overline{V|_{M_P}}$.
3040: \label{itemSplitVanishingPTauInvariant}
3041: \end{enumerate}
3042: If either condition holds, $V= H^{\l(w)}(\n_P;E)_w$ for $w\in W_P$ a
3043: fundamental Weyl element for $P$ in $G$ and
3044: \itemref{itemSplitVanishingPTauInvariant} holds for any such $V$.
3045: \end{lem}
3046: 
3047: \begin{proof}
3048: Let $\h$ be a Cartan subalgebra $\h$ of $\levi_P$ and fix positive
3049: orderings for the roots of $\levi_{P\CC}$ and $\mathfrak g_\CC$ as in Lemma
3050: ~\ref{ssectFundamentalParabolic}.  Let $E$ have highest weight $\lambda$;
3051: by Kostant's theorem, for all $w\in W_P$ the $L_P$-module
3052: $V=H^{\l(w)}(\n_P;E)_w$ has highest weight $\u=w(\lambda+\hsr)-\hsr$.  One
3053: may use \S\S\ref{ssectConjugateSelfContragredient} and
3054: \ref{ssectConjugateSelfContragredientExample} to translate
3055: \itemref{itemFundamentalGTauInvariant} and
3056: \itemref{itemSplitVanishingPTauInvariant} into assertions regarding $w$ and
3057: $\lambda$; their equivalence in this form is proved in
3058: \cite[~3.6(iii)(iv)]{refnBorelCasselman} (one must replace $\lambda$ in
3059: \cite{refnBorelCasselman} by $\lambda+\hsr$ to agree with our notation).
3060: For the final assertion, note that since $\t_P$ preserves the decomposition
3061: $\hb_G = \hb_P + \sa^G_P$ and acts as $-1$ on the second factor,
3062: \itemref{itemSplitVanishingPTauInvariant} holds for a given $w$ if and only
3063: if $\t_Pw(\lambda+\hsr)|_{\hb_G}=w(\lambda+\hsr)|_{\hb_G}$ and hence if and
3064: only if $w^{-1}\t_Pw((\lambda+\hsr)|_{\hb_G})=(\lambda+\hsr)|_{\hb_G}$.
3065: Since $w^{-1}\t_Pw$ and $\t_G$ are members of $W$ after composing with $-c$
3066: and they agree on a regular element, they must be equal.  (In fact this
3067: argument may be completed to directly prove the equivalence of
3068: \itemref{itemFundamentalGTauInvariant} and
3069: \itemref{itemSplitVanishingPTauInvariant}.)
3070: \end{proof}
3071: 
3072: \section{Alternate Forms of Definition
3073: ~\ref{ssectMicroSupport}\itemref{itemMicroSupportQType}}
3074: \label{sectAlternateMicroSupport}
3075: Let $W$ be an admissible space possessing a unique maximal stratum $X_R$.
3076: 
3077: \subsection{A partial ordering on $\IrrRep(\L_W)$}
3078: \label{ssectPartialOrdering}
3079: Let $V$, $\tilde V\in \IrrRep(\L_W)$ be irreducible $L_P$- and
3080: $L_{\tilde P}$-modules respectively.  Set $V \preccurlyeq_0 \tilde V$ if
3081: the following three conditions are fulfilled:
3082: \begin{enumerate}
3083: \item $P\le \tilde P$;
3084: \item \label{itemZeroKostantComponent} $V= H^{\l(w)}(\n_{P}^{\tilde P};
3085: \tilde V)_w$ for an element $w\in W_P^{\tilde P}$;
3086: %\item $V= H^{\l(w)}(\n_{P}^{\tilde P}; \tilde V)_w$ for a fundamental Weyl
3087: %element $w\in W_P^{\tilde P}$;
3088: \item \label{itemZero} $(\xi_V+\hsr)|_{\sa_P^{\tilde P}} = 0$.
3089: \end{enumerate}
3090: From \S\ref{ssectKostantDegeneration} it is easy to see that this is a
3091: partial order.  If $V\preccurlyeq_0 \tilde V$ we set $[\tilde V:V]=\l(w)$,
3092: where $w$ is as in \itemref{itemZeroKostantComponent}.  (We will consider
3093: the partial ordering obtained by relaxing \itemref{itemZero} later in
3094: \S\ref{ssectAdditionalPartialOrdering}.)
3095: 
3096: Lemma ~\ref{ssectBorelCasselman} immediately implies the following
3097: 
3098: \begin{lem*}
3099: If $V \preccurlyeq_0 \tilde V$ and $(V|_{M_P})^*\cong \overline{V|_{M_P}}$
3100: then $ (\tilde V|_{M_{\tilde P}})^*\cong \overline{\tilde V|_{M_{\tilde
3101: P}}}$, the Weyl element $w$ from \itemref{itemZeroKostantComponent} above
3102: is fundamental for $P$ in $\tilde P$, and $[\tilde V: V] = \tfrac12 \dim
3103: \n_P^{\tilde P}$.
3104: \end{lem*}
3105: 
3106: \begin{prop}
3107: \label{ssectAlternateMicroSupport}
3108: Let $\M$ be an $\L_W$-module and let $V\in \IrrRep(\L_W)$ be an irreducible
3109: $L_P$-module.  The following conditions on $V$ are equivalent\textup:
3110: \begin{enumerate}
3111: \item $H(\i_{P}^* \ihat_{Q}^! \M)_V \neq 0$ for some $Q\in [Q_V^W,
3112: Q^{\prime W}_V]$.
3113: \label{itemSSCondition}
3114: \item $H(\i_P^* \i_{{\tilde P}*}\i_{\tilde P}^* \ihat_{\tilde Q}^!\M)_V
3115: \neq 0$ for some $\tilde Q \in Q_V^W, Q^{\prime W}_V]$ and $\tilde P =
3116: (P,Q_V^W)\cap \tilde Q$.
3117: \label{itemFaryTermCondition}
3118: \item
3119: \label{itemBasicSSCondition}
3120: There exists an $L_{\tilde P}$-module $\tilde V\succcurlyeq_0 V$ such that
3121: $H(\i_{\tilde P}^* \ihat_{{Q^W_{\tilde V}}}^!  \M)_{\tilde V} \neq 0$.
3122: \item
3123: \label{itemessSSCondition}
3124: There exists an $L_{\tilde{\tilde P}}$-module $\tilde{\tilde V}\succcurlyeq_0
3125: V$ such that
3126: \begin{equation*}
3127: \Im\bigl ( H(\i_{\tilde{\tilde P}}^* \ihat_{{Q^W_{\tilde{\tilde V}}}}^!
3128: \M)_{\tilde{\tilde V}} \longrightarrow H(\i_{\tilde{\tilde P}}^*
3129: \ihat_{Q^{\prime W}_{\tilde{\tilde V}}}^!  \M)_{\tilde{\tilde V}} \bigr)
3130: \neq 0\ .
3131: \end{equation*}
3132: \end{enumerate}
3133: If these conditions hold it can be arranged that $\tilde P$ in
3134: \itemref{itemFaryTermCondition} and
3135: \itemref{itemBasicSSCondition} is the same and that
3136: $\tilde{\tilde P}\ge \tilde P$.  Furthermore, let
3137: $[c_i(\cdot;\M),d_i(\cdot;\M)]$ \textup($i=1$, \dots, $4$\textup)
3138: denote the range of degrees in which the expressions in
3139: \itemref{itemSSCondition}--\itemref{itemessSSCondition}
3140: respectively can be nonvanishing, where in the parentheses we indicate
3141: the module and, if desired, the parabolic $\QQ$-subgroup to which we wish to
3142: restrict.  Then
3143: \begin{equation}
3144: \label{eqnAlternateDegreeRanges}
3145: \begin{aligned}
3146: {[c_1(V;\M),d_1(V;\M)]} &\subseteq [c_2(V;\M), d_2(V;\M)] \ ,\\ 
3147: [c_2(V,\tilde P;\M),d_2(V,\tilde P;\M)] &= [c_3(\tilde V;\M),d_3(\tilde V;\M)] +
3148: \tfrac12\dim \n_P^{\tilde P} \ ,\\
3149: [c_3(\tilde V;\M),d_3(\tilde V;\M)] &\subseteq [c_4(\tilde{\tilde
3150: V};\M),d_4(\tilde{\tilde V};\M) + \dim \sa_{\tilde P}^{\tilde{\tilde P}}] +
3151: \tfrac12\dim \n_{\tilde P}^{\tilde{\tilde P}} \ .
3152: \end{aligned}
3153: \end{equation}
3154: \end{prop}
3155: 
3156: \begin{proof}
3157: \itemref{itemSSCondition}$\Longleftrightarrow$\itemref{itemFaryTermCondition}:
3158: Assume \itemref{itemSSCondition} holds and consider the long exact sequence
3159: \begin{equation}
3160: \cdots \longrightarrow
3161: H^j(\i_P^* \ihat_{Q_V^W}^!  \M)_V \xrightarrow{\b}
3162: H^j(\i_P^* \ihat_{Q}^! \M)_V \longrightarrow
3163: H^{j}(\i_P^* \jhat_{Q_V^W*}\jhat_{Q_V^W}^*\ihat_{Q}^! \M)_V \longrightarrow
3164: \cdots
3165: \label{eqnCompareTypes}
3166: \end{equation}
3167: as in \eqref{eqnLongCompareLocalCohomologyWithSupports}.  If $\b$ is
3168: surjective, then $H(\i_{ P}^* \ihat_{{Q_V^W}}^!  \M)_{ V}\neq 0$ and
3169: \itemref{itemFaryTermCondition} holds with $\tilde Q = Q_V^W$ and
3170: $\tilde P = P$.  Otherwise the last term of \eqref{eqnCompareTypes} is
3171: nonzero which implies by the Fary spectral sequence of
3172: Lemma~\ref{ssectRelativeLocalCohomologySupportsSS} and equation
3173: \eqref{eqnPushForwardType} that \itemref{itemFaryTermCondition} holds
3174: for some $\tilde Q\le Q$ with $\tilde P > P$.  Conversely suppose
3175: \itemref{itemFaryTermCondition} holds for some $\tilde Q$ and consider
3176: the analogous long exact sequence to \eqref{eqnCompareTypes} in which
3177: $Q$ has been replaced by $\tilde Q$.  By
3178: \itemref{itemFaryTermCondition} the Fary spectral sequence for the
3179: last term has $E_1$ nonzero at the top level.  Thus either the last
3180: term is nonzero, in which case \itemref{itemSSCondition} holds for
3181: $Q_V^W$ or $\tilde Q$, or else the last term is zero, in which case
3182: \itemref{itemFaryTermCondition} holds also for some smaller $\tilde Q$
3183: and we may repeat the argument.
3184: 
3185: \itemref{itemFaryTermCondition}$\Longleftrightarrow$\itemref{itemBasicSSCondition}:
3186: Assume that \itemref{itemFaryTermCondition} holds.  Then by
3187: \eqref{eqnPushForwardType} we see that $H(\i_{\tilde P}^*\ihat_{\tilde
3188: Q}^!\M)_{\tilde V}\neq 0$ for some irreducible $L_{\tilde P}$-module
3189: $\tilde V$ such that $V$ is an irreducible constituent of $H(\n_{P}^{\tilde
3190: P}; \tilde V)$.  By the definition of $\tilde P$ we see that
3191: $(\xi_V+\hsr)|_{\sa_P^{\tilde P}}=0$ and so $V\preccurlyeq_0 \tilde V$.
3192: Finally $(\xi_V+\hsr)|_{\sa_{\tilde P}} = \xi_{\tilde V}+\hsr_{\tilde P}$
3193: implies that $\tilde Q = Q^W_{\tilde V}$ and thus
3194: \itemref{itemBasicSSCondition} holds.  Conversely
3195: \itemref{itemBasicSSCondition} implies $H(\i_P^* \i_{{\tilde P}*}\i_{\tilde
3196: P}^* \ihat_{Q^W_{\tilde V}}^!\M)_V \neq 0$.  We have $Q^W_{\tilde V} =
3197: Q_V^W\vee \tilde P$ and $Q^{\prime W}_{\tilde V} = Q^{\prime W}_V$ so $
3198: Q^W_{\tilde V} \in [Q_V^W, Q^{\prime W}_V]$ and hence
3199: \itemref{itemFaryTermCondition} holds.
3200: 
3201: \itemref{itemBasicSSCondition}$\Longleftrightarrow$\itemref{itemessSSCondition}:
3202: Assume that \itemref{itemBasicSSCondition} holds and consider the long
3203: exact sequence
3204: \begin{equation*}
3205: \cdots \longrightarrow
3206: H^{j-1}(\i_{\tilde P}^* \jhat_{Q^W_{\tilde V}*}\jhat_{Q^W_{\tilde
3207:     V}}^*\ihat_{Q^{\prime W}_{\tilde V}}^! \M)_{\tilde V} \longrightarrow
3208: H^j(\i_{\tilde P}^* \ihat_{Q^W_{\tilde V}}^!  \M)_{\tilde V} \xrightarrow{\b}
3209: H^j(\i_{\tilde P}^* \ihat_{Q^{\prime W}_{\tilde V}}^! \M)_{\tilde V} \longrightarrow
3210: \cdots \ .
3211: \end{equation*}
3212: Either $\b$ is nonzero, in which case \itemref{itemessSSCondition} holds,
3213: or the first term of the sequence is nonzero.  By
3214: Lemma~\ref{ssectRelativeLocalCohomologySupportsSS} and equation
3215: \eqref{eqnPushForwardType} this implies that
3216: \itemref{itemFaryTermCondition} and hence \itemref{itemBasicSSCondition}
3217: holds for a larger $\tilde P$ and we may repeat the argument.  Obviously
3218: \itemref{itemessSSCondition}$\implies$\itemref{itemBasicSSCondition}.
3219: \end{proof}
3220: 
3221: \begin{cor}
3222: \label{ssectPartialOrderingOnMicroSupport}
3223: If $V\in \mS(\M)$ there exists $\tilde V\in \emS(\M)$ with $V\preccurlyeq_0
3224: \tilde V$.  If $\tilde V \in \mS(\M)$ and $V\in\IrrRep(\L_W)$ is an
3225: irreducible $L_P$-module for which $V|_{M_P}$ is conjugate
3226: self-contragredient and $V\preccurlyeq_0 \tilde V$, then $V\in \mS(\M)$.
3227: \end{cor}
3228: 
3229: Thus $\emS(\M)$ determines $\mS(\M)$; the converse is not true
3230: however---this is because the implication
3231: \itemref{itemFaryTermCondition}$\implies$\itemref{itemSSCondition}
3232: does not determine $Q$.
3233: 
3234: \specialsection*{Part II. A Global Vanishing Theorem for $\L$-modules}
3235: In this part we present a vanishing theorem for the global cohomology of an
3236: $\L$-module $\M$.  The well-known local-global principle implies that if
3237: the local cohomology of a sheaf vanishes, then the global hypercohomology
3238: vanishes.  Our result implies that if the micro-support of an $\L$-module
3239: is empty, then the global cohomology vanishes.  More generally we give a
3240: bound based on the micro-support on the degrees in which cohomology can be
3241: nonzero.
3242: 
3243: \section{The Vanishing Theorem}
3244: \label{sectVanishingTheoremLModule}
3245: 
3246: \subsection{}
3247: \label{ssectWeightCentralizer}
3248: Let $P\in\Pl$ and let $\h$ be a Cartan subalgebra of $\levi_P$.  Given
3249: $\u\in\h_\CC^*$ view $\u$ as an element of $\levi_{P\CC}^*$ by extending
3250: $\u$ to be zero on all root spaces.  Let $L_P(\u)\subseteq L_P$ be the
3251: stabilizer of $\u$ under the coadjoint action.  This is a reductive
3252: subgroup (defined over $\CC$) whose Lie algebra $\levi_P(\u)$ is generated
3253: by $\h$ and the root spaces for
3254: \begin{equation}
3255: \{\,\g\in\Phi(\levi_{P\CC},\h_{\CC}) \mid
3256:          \langle\u,\g\spcheck\rangle=0\,\}\ .
3257: \label{eqnLPxiRoots}
3258: \end{equation}
3259: As extreme cases we have $L_P(0)=L_P$ and $L_P(\hsr)=H$, the Cartan
3260: subgroup.
3261: 
3262: \subsection{}
3263: \label{ssectVcentralizer}
3264: Let $V$ be an irreducible regular $L_P$-module for which $(V|_{M_P})^*\cong
3265: \overline{V|_{M_P}}$.  Choose $\h$ and $\Phi^+$ so that $\t_P=\theta$ for
3266: some Cartan involution and let $\u\in \h^*_\CC$ be the corresponding
3267: highest weight of $V$.  In this case $\overline{\u|_{\hb_{P}}} =
3268: -\u|_{\hb_{P}}$ and so $L_P(\u)$ is defined over $\RR$.  Let
3269: \begin{equation*}
3270: D_P(\u) = L_P(\u)(\RR)/(K_P\cap L_P(\u))A_P
3271: \end{equation*}
3272: denote the corresponding symmetric space (compare
3273: \cite{refnBorelVanishingTheorem}).  The dimension of $D_P(\u)$ can vary
3274: depending upon the $\theta$-stable positive system $\Phi^+$; we let
3275: $D_P(V)$ denote any one of the $D_P(\u)$ with maximal dimension.  Then
3276: $\dim D_P(V)$ is well-defined and independent of the choice of $\h$ and
3277: $\theta$.  It will be convenient beginning in
3278: \S\ref{sectEqualRankBasicLemma} to allow $V$ in $D_P(V)$ to be merely
3279: isotypical as opposed to irreducible.
3280: 
3281: \subsection{}
3282: \label{ssectcMdMDefinition}
3283: Let $\M$ be an $\L$-module on $\Xhat$.  For $V\in\mS(\M)$ let $c(V;\M)\le
3284: d(V;\M)$ be the least and greatest degrees in which $H^i(\i_{P}^*
3285: \ihat_{Q}^! \M)_V \neq 0$ (for any $Q\in[ Q_V,Q_V']$).  Define
3286: \begin{align}
3287: \label{eqncMDefn}
3288: c(\M) &= \inf_{V\in\mS(\M)} \tfrac12(\dim D_P - \dim D_P(V)) + c(V;\M)\ ,
3289: \text{ and} \\
3290: \label{eqndMDefn}
3291: d(\M) &= \sup_{V\in\mS(\M)} \tfrac12(\dim D_P + \dim D_P(V)) + d(V;\M)\ .
3292: \end{align}
3293: Note that the shift $\tfrac12(\dim D_P \pm \dim D_P(V))$ corresponds to the
3294: range of degrees in which $H_{(2)}(X_P;\VV)$ can be nonzero (see Theorem
3295: ~\ref{ssectRaghunathanVanishing}).
3296: 
3297: \begin{thm}
3298: \label{ssectGlobalVanishing}
3299: Let $\M$ be an $\L$-module on $\Xhat$. Then  $H^i(\Xhat;\M)=0$ for $i
3300: \notin [c(\M), d(\M)]$.
3301: \end{thm}
3302: 
3303: The proof will appear in \S\ref{sectProofGlobalVanishing} after a number of
3304: analytic preliminaries.  In the remainder of this section we will indicate
3305: how $c(\M)$ and $d(\M)$ may be computed in terms of the essential
3306: micro-support.
3307: 
3308: \begin{lem}
3309: \label{ssectDPVInequality}
3310: Let $V$ be an irreducible regular $L_P$-module for $P\in\Pl$ and assume
3311: $V|_{M_P}$ is conjugate self-contragredient.  If $V\preccurlyeq_0
3312: \tilde V$ then
3313: \begin{equation}
3314: \label{eqnComparedimDPV}
3315: \dim D_P(V) \le \dim D_{\tilde P}(\tilde V) - \dim
3316: \sa_P^{\tilde P}\ .
3317: \end{equation}
3318: Consequently
3319: \begin{align}
3320: \label{eqnLowerdimDPInequality}
3321: \tfrac12\bigl(\dim D_P - \dim D_P(V)\bigr) &\ge \tfrac12\bigl(\dim
3322: D_{\tilde P} - \dim D_{\tilde P}(\tilde V) \bigr) -\tfrac12\dim
3323: \n_{P}^{{\tilde P}}\ , \\
3324: \label{eqnUpperdimDPInequality}
3325: \tfrac12\bigl(\dim D_P + \dim D_P(V)\bigr) &\le \tfrac12\bigl(\dim
3326: D_{{\tilde P}} + \dim D_{{\tilde P}}({\tilde V}) \bigr)  -\tfrac12\dim
3327: \n_{P}^{{\tilde P}} \\
3328: &\qquad - \dim \sa_P^{{\tilde P}}\ . \notag
3329: \end{align}
3330: \end{lem}
3331: 
3332: \begin{proof}
3333: Let $\h$ be a $\theta$-stable Cartan subalgebra for $\levi_P$ and hence for
3334: $\levi_{\tilde P}$.  Choose $\h$ and $\Phi^+$ so that $\t_P=\theta$ and let
3335: $\u$ be the corresponding highest weight of $V$.  Then we can define
3336: $D_P(\u)$ to be the symmetric space of $L_P(\u)(\RR)$ as in
3337: \S\ref{ssectVcentralizer}.  Now by hypothesis and Lemma
3338: ~\ref{ssectBorelCasselman} we know that $V=H^{\l(w)}(\n_P^{\tilde P};\tilde
3339: V)_w$ for $w\in W_P^{\tilde P}$ such that $w\t_{\tilde P}w^{-1} = \t_P$.
3340: Then $\widetilde\t_{\tilde P}=\theta$, where $\widetilde\t_{\tilde P}=
3341: w\t_{\tilde P}w^{-1}$ is the operator of \S\ref{ssectDefineTau} defined
3342: with respect to the positive system $w\Phi^+$.  So if $\tilde\u$ is the
3343: highest weight of $\tilde V$ with respect to $\Phi^+$, we can define
3344: $D_{\tilde P}(w\tilde\u)$ to be the symmetric space of $L_{\tilde
3345: P}(w\tilde\u)(\RR)$ as in \S\ref{ssectVcentralizer}.
3346: 
3347: Now $\u=w(\tilde\u+\hsr)-\hsr$ by Kostant's theorem
3348: \S\ref{ssectKostantsTheorem}.  So if $\g$ is a root of $\levi_{P\CC}$ and
3349: $\langle\u,\g\spcheck\rangle=0$ then $\langle
3350: w\tilde\u,\g\spcheck\rangle=0$ as well, since both $w\tilde\u$ and
3351: $w\hsr-\hsr$ are dominant for $\Phi^+(\levi_{P\CC},\h_\CC)$.  Thus
3352: $L_P(\u)\subseteq L_{\tilde P}(w\tilde\u)$ and hence $\dim D_P(\u) \le \dim
3353: D_{\tilde P}(w\tilde\u) - \dim \sa_P^{\tilde P}$; the subtraction of $\dim
3354: \sa_P^{\tilde P}$ adjusts for taking quotient by $A_P$ on the left and
3355: $A_{\tilde P}$ on the right.  The lemma follows by taking the maximum over
3356: possible $\Phi^+$ and by applying $\dim D_P = \dim D_{\tilde P} - \dim
3357: \sa_P^{\tilde P} - \dim \n_P^{\tilde P}$.
3358: \end{proof}
3359: 
3360: \subsection{}
3361: \label{ssectEqualityDegreeRanges}
3362: Let $\M$ be an $\L$-module on $\Xhat$.  For $V\in\mS(\M)$ let
3363: $[c_i(V;\M),d_i(V;\M)]$ ($i=1$, \dots, $4$) be as in Proposition
3364: ~\ref{ssectAlternateMicroSupport}.  Define $c_i(\M)$ and
3365: $d_i(\M)$ analogously to \eqref{eqncMDefn} and \eqref{eqndMDefn} with the
3366: exception that we set 
3367: \begin{equation*}
3368: d_2(\M) = \sup_{V\in\mS(\M),\tilde P} \tfrac12(\dim D_P + \dim D_P(V))
3369: + \dim \sa_P^{\tilde P} + d_2(V,\tilde P;\M)\ .
3370: \end{equation*}
3371: Of course $[c(\M),d(\M)]=[c_1(\M),d_1(\M)]$.  In fact we have
3372: 
3373: \begin{prop*}
3374: The intervals $[c_i(\M),d_i(\M)]$ are equal for $i=1$, \dots, $4$.
3375: \end{prop*}
3376: 
3377: \begin{proof}
3378: We will prove all $d_i(\M)$ are equal; the proof of the other
3379: equalities is similar.  For $V\in\mS(\M)$ let $V\preccurlyeq_0 \tilde V
3380: \preccurlyeq_0 \tilde{\tilde V}$ be chosen as in Proposition
3381: ~\ref{ssectAlternateMicroSupport} such that $d_2(V;\M) = d_2(V,\tilde
3382: P;\M)$.  Then
3383: \begin{equation*}
3384: \begin{split}
3385: d_1(V;\M) \le  d_2(V,\tilde P;\M) &= d_3(\tilde V;\M) +
3386: \tfrac12\dim\n_P^{\tilde P} \\
3387: &\le d_4(\tilde{\tilde V};\M) + \dim \sa_{\tilde
3388: P}^{\tilde{\tilde P}} + \tfrac12\dim \n_{P}^{\tilde{\tilde P}}
3389: \end{split}
3390: \end{equation*}
3391: by \eqref{eqnAlternateDegreeRanges}.  These inequalities together with
3392: \eqref{eqnUpperdimDPInequality} (applied both to $V\preccurlyeq_0 \tilde V$
3393: and $\tilde V\preccurlyeq_0 \tilde{\tilde V}$) yield $d_1(\M)\le d_2(\M)\le
3394: d_3(\M) \le d_4(\M)$.  On the other hand, for $V\in \emS(\M)$ the obvious
3395: inequality $d_4(V;\M) \le d_1(V;\M)$ implies that $d_4(\M)\le d_1(\M)$.
3396: \end{proof}
3397: 
3398: \begin{cor*}
3399: For $V\in\emS(\M)$ let $\essc(V;\M)\le
3400: \essd(V;\M)$ be the least and greatest degrees in which
3401: \begin{equation*}
3402: \Im \bigl(H(\i_{P}^* \ihat_{{Q_V^W}}^! \M)_V \longrightarrow H(\i_{P}^*
3403: \ihat_{Q^{\prime W}_V}^! \M)_V \bigr) \neq 0 \ .
3404: \end{equation*}
3405: Then
3406: \begin{align*}
3407: c(\M) &= \inf_{V\in\emS(\M)} \tfrac12(\dim D_P - \dim D_P(V)) + \essc(V;\M)\ ,
3408: \text{ and} \\
3409: d(\M) &= \sup_{V\in\emS(\M)} \tfrac12(\dim D_P + \dim D_P(V)) + \essd(V;\M)\ .
3410: \end{align*}
3411: \end{cor*}
3412: 
3413: \section{$L^2$-cohomology with Boundary Values}
3414: \label{sectLtwo}
3415: In this section we introduce a variant of $L^2$-cohomology whose elements
3416: admit restrictions to boundary faces.
3417: 
3418: \subsection{$L^2$-cohomology}
3419: \label{ssectLtwoCohomology}
3420: We first recall ordinary $L^2$-cohomology; references are
3421: \cite{refnCheeger} and \cite{refnZuckerWarped}.  Let $Y$ be a Riemannian
3422: manifold (possibly with boundary or corners) and let $\EE$ be a metrized
3423: locally constant sheaf.  For a measurable $\EE$-valued form $\o$ on $Y$ let
3424: \begin{equation*}
3425: \lVert\o\rVert=\lVert\o\rVert_Y=\left(\int_Y \lvert\o\rvert^2
3426:  \,dV\right)^{\frac12}
3427: \end{equation*}
3428: denote the usual $L^2$ norm (we will omit the subscript $Y$ if this would
3429: not cause confusion).    Let $L_2(Y;\EE)$ be the Hilbert space of those
3430: forms $\o$ for which $\lVert\o\rVert < \infty$; equivalently this is the
3431: completion of $A_c(Y;\EE)$ with respect to the $L^2$ norm.  Let
3432: $d=d_Y=d_{Y,\EE}$ denote the unbounded densely defined operator given by
3433: exterior differentiation on the domain
3434: \begin{equation*}
3435: \dom d = \{\,\o\in A(Y;\EE)\mid
3436: \lVert\o\rVert,\lVert d\o\rVert<\infty\,\}
3437: \end{equation*}
3438: and let $\dbar$ be its closure in the graph norm.  That is, $\o\in
3439: L_{2}(Y;\EE)$ is in $\dom \dbar$ if and only if there exists a
3440: sequence $\o_i\in\dom d$ such $\o_i$ and $d\o_i$ are Cauchy sequences and
3441: $\o_i \to \o$ in $L^2$ norm.  The {\itshape $L^2$-cohomology\/}
3442: $H_{(2)}(Y;\EE)$ is defined to be the cohomology of the complex $\dom
3443: \dbar$.  A smoothing argument \cite[\S15]{refndeRham},
3444: \cite[\S8]{refnCheeger} shows that the inclusion $\dom d\subseteq \dom
3445: \dbar$ induces an isomorphism $H_{(2)}(Y;\EE)\cong H(\dom d)$.
3446: Consequently
3447: \begin{equation}
3448: \label{eqnLtwocohomologyIsOrdinaryForCompact}
3449: H_{(2)}(Y;\EE)\cong H(Y;\EE) \qquad\text{if $Y$ is compact.}
3450: \end{equation}
3451: 
3452: We give $H_{(2)}(Y;\EE)$ the quotient topology; this is Hausdorff if and
3453: only if $\Range \dbar$ is closed.  If it is Hausdorff, then
3454: $H_{(2)}(Y;\EE)$ is representable by harmonic forms, namely $\ker \dbar
3455: \cap \ker \dbar^*$ where $\dbar^*$ is the adjoint of $\dbar$ in the sense
3456: of unbounded operators on the Hilbert space $L_2(Y;\EE)$; conversely if it
3457: is not Hausdorff the $L^2$-cohomology is infinite dimensional.
3458: 
3459: If $g>0$ is a positive function on $Y$ one may also consider the weighted
3460: norm $\lVert\o\rVert_g^2 = \int_Y \lvert\o\rvert^2 \,gdV$, the
3461: corresponding Hilbert space $L_2(Y;\EE,g)$, and the $L^2$-cohomology
3462: $H_{(2)}(Y;\EE,g)$.  Since this case may be absorbed into the previous by
3463: modifying the metric on $\EE$, we do not mention it again in this section.
3464: 
3465: \subsection{Spectral vanishing criterion for $H_{(2)}(Y;\EE)$}
3466: \label{ssectSpectralVanishingLtwoCohomology}
3467: If $Y$ is complete as a metric space, a certain estimate implies the
3468: vanishing of $L^2$-cohomology.  More precisely, define the space of
3469: \emph{$\epsilon$-approximate harmonic forms}
3470: \begin{equation*}
3471: \H_\epsilon(Y;\EE)=
3472: \{\,\o\in \dom \dbar^* \cap
3473: A_c(Y;\EE) \mid \lVert d \o \rVert^2 + \lVert d^*\o\rVert^2 \le \epsilon
3474: \lVert \o \rVert^2 \,\}\ .
3475: \end{equation*}
3476: (Note that the intersection with $\dom \dbar^*$, which forces $\o$ to
3477: satisfy Neumann boundary conditions, may be omitted if $\partial
3478: Y=\emptyset$.)
3479: 
3480: \begin{prop*}
3481: \textup(\cite[Prop.~1.2]{refnSaperSternTwo}\textup) Assume that $Y$ is
3482: complete as a metric space.  Then $\H_{\epsilon}^i(Y;\EE) = 0$ for some
3483: $\epsilon>0$ if and only if $H_{(2)}^i(Y;\EE)=0$ and $H_{(2)}^{i+1}(Y;\EE)$
3484: is Hausdorff.
3485: \end{prop*}
3486: 
3487: \subsection{$L^2$-cohomology with boundary values}
3488: \label{ssectLtwoBoundaryValues}
3489: For $\o\in A(Y;\EE)$ define the norm
3490: \begin{equation*}
3491: \lvvv\o\rvvv^2 = \|\o\|_Y^2 + \sum_F \| \o|_{F} \|_F^2,
3492: \end{equation*}
3493: where $F$ ranges over the closed proper boundary faces of $Y$.  We will
3494: write, for example, ``$\lvvv\cdot\rvvv$-bounded'' if we mean to use this
3495: norm.  Let $d_b= d_{Y,b}=d_{Y,b,\EE}$ denote the unbounded operator on
3496: $L_{2}(Y;\EE)$ given by exterior differentiation on the domain
3497: \begin{equation*}
3498: \Dom d_{b} = \{\,\o\in A(Y;\EE) \mid \lvvv\o\rvvv, \lvvv
3499: d\o\rvvv<\infty\,\}
3500: \end{equation*}
3501: and let $\dbar_{b}$ be its $\lvvv\cdot\rvvv$-closure in the graph norm.
3502: That is, $\o\in L_{2}(Y;\EE)$ is in $\dom \dbar_b$ if and only if
3503: there exists a sequence $\o_i\in\dom d_b$ such $\o_i$ and $d\o_i$ are
3504: Cauchy sequences in the norm $\lvvv\cdot\rvvv$ and $\| \o_i - \o\|\to 0$
3505: (usual $L^2$ norm).  Since $\left\|\cdot\right \| \le \lvvv\cdot\rvvv$, we
3506: have an inclusion of complexes $\dom \dbar_b \subseteq \dom \dbar$, but it
3507: is important to note that $\dbar_b$ is not a closed operator on
3508: $L_{2}(Y;\EE)$ in general.  (In \S\ref{ssectLtwoBoundarySpace} below
3509: we will consider a larger Hilbert space in which $\dbar_b$ is closed.)  The
3510: {\itshape $L^2$-cohomology with boundary values\/} is defined to be
3511: \begin{equation*}
3512: H_{(2),b}(Y;\EE) = H(\Dom\dbar_b)\ .
3513: \end{equation*}
3514: (Actually in general one would want to consider various weight functions on
3515: the boundary faces in defining $\lvvv\cdot\rvvv$, but this is not
3516: necessary here; see Remark ~\ref{ssectGeneralBoundaryNorm} below.)
3517: 
3518: \subsection{}
3519: $L^2$-cohomology with boundary values is convenient since $\o\in \dom
3520: \dbar_b$ admits a strong $L_2$ trace or restriction $\o|_F\in
3521: L_2(F;\EE|_F)$ for each closed boundary face $F$.  Specifically
3522: 
3523: \begin{lem*}
3524: Restriction to a closed bounded face $F$ induces a well-defined
3525: $\lvvv\cdot\rvvv$-bounded map
3526: \begin{equation*}
3527: \Dom\dbar_{Y,b}\to \Dom\dbar_{F,b}
3528: \end{equation*}
3529: and consequently a map
3530: \begin{equation*}
3531: H_{(2),b}(Y;\EE) \longrightarrow H_{(2),b}(F;\EE|_F) \ .
3532: \end{equation*}
3533: \end{lem*}
3534: 
3535: \begin{proof}
3536: Given $\o\in \dom \dbar_b$ let $\o_i$ be a sequence of smooth forms with
3537: $\lvvv \o_i - \o\rvvv\to 0$ and $d\o_i$ a Cauchy sequence in the norm
3538: $\lvvv\cdot\rvvv$.  Then $\o_i|_F $ is a Cauchy sequence in
3539: $L_2(F;\EE|_F)$ and we set $\o|_F$ to be its limit.  Clearly
3540: $\o|_F\in \dom \dbar_{F,b}$.
3541: 
3542: To see that $\o|_F$ is well-defined, it suffices to consider $\o = 0$
3543: and $F$ a codimension 1 boundary face.  (For if $\o|_F$ is
3544: well-defined and $F'\subset F$, then $\o|_F\in \dom\dbar_{F,b}$ and
3545: $\o|_{F'} = (\o|_F)|_{F'}$.)  Let $\Int F$ denote the interior of $F$
3546: and $\v$ the outward unit normal vector.  For any form $\al\in
3547: A_c(\Int F;\EE|_F)$, there clearly exists a form $\tilde \al\in
3548: A_c(Y;\EE)$ of one degree higher with Neumann boundary data $(\iota_{\v}\tilde
3549: \al)|_F = \al$.  Then Stokes's theorem implies that
3550: \begin{equation*}
3551: (\o_i|_F,\al)_F = (d\o_i,\atilde) - (\o_i,\CoDifferential \atilde)
3552: \end{equation*}
3553: where $\CoDifferential$ is the formal adjoint to $d$.  Now if $\o_i
3554: \to 0$ with $d\o_i$ Cauchy, we would have $d\o_i \to 0$ as usual and
3555: hence by the above formula, $ (\o_i|_F,\al)_F\to 0$ for all $\al$.
3556: Since such $\al$ are dense in $L_2(F;\EE|_F)$, we see that
3557: $\o|_F = \lim \o_i|_F = 0$.
3558: \end{proof}
3559: 
3560: \subsection{}
3561: \label{ssectBoundaryAveraging}
3562: On the other hand, we have an inclusion of complexes $\Dom\dbar_b
3563: \hookrightarrow \Dom\dbar$ which induces a homomorphism
3564: \begin{equation*}
3565: H_{(2),b}(Y;\EE) \longrightarrow H_{(2)}(Y;\EE).
3566: \end{equation*}
3567: We will give a simple criterion for this to be an isomorphism which will
3568: suffice for our purposes.
3569: 
3570: Consider a system of tubular neighborhoods
3571: \begin{equation*}
3572: (N_F, p_F, t_F)
3573: \end{equation*}
3574: of the closed boundary faces $F\subset Y$, where $N_F$ is an open
3575: neighborhood of $F$, $p_F\colon N_F \to F$ is a smooth retraction, $t_F\colon N_F \to
3576: [0,1)^{\codim F}$ is a normal variable with $F=t_F^{-1}(0)$, and
3577: \begin{equation*}
3578: (t_F,p_F)\colon  N_F \tildearrow [0,1)^{\codim F}\times F
3579: \end{equation*}
3580: is a smooth diffeomorphism.  We require that this data is compatible in the
3581: sense that if $F$ and $F'$ intersect, then
3582: \begin{equation}
3583: \begin{aligned}
3584: \relax&N_{F\cap F'}=N_F\cap N_{F'},\\
3585: \relax&p_{F\cap F'} = p_Fp_{F'} = p_{F'}p_F \qquad\text{on $N_{F\cap F'}$},\\
3586: \relax&t_{F\cap F'} = (t_F, t_{F'})\qquad\text{on $N_{F\cap F'}$, up to permutation
3587: of coordinates.}
3588: \end{aligned}
3589: \label{eqnNeighborhoodCompatibility}
3590: \end{equation}
3591: It is not difficult to see that for any manifold with corners (always
3592: assumed paracompact) such a system of tubular neighborhoods always exists.
3593: Note also that such data induces canonical isomorphisms
3594: \begin{equation*}
3595: \EE|_{N_F} \cong p_F^* \EE|_F.
3596: \end{equation*}
3597: 
3598: \begin{lem*}
3599: Let $Y$ be a Riemannian
3600: manifold with corners and $\EE$ a metrized locally constant sheaf on $Y$.
3601: Assume that a system of tubular neighborhoods satisfying
3602: \eqref{eqnNeighborhoodCompatibility} can be chosen so that the
3603: Riemannian metric $ds^2$ on $Y$ and the metric $h$ on $\EE$ satisfy the
3604: quasi-isometry%
3605: %
3606: \footnote{Two metrics $g$, $g'$ are said to be {\textit quasi-isometric\/}
3607: if there exist constants $C$, $C'>0$ such that $Cg'\le g \le Cg'$; we
3608: denote this by $g\sim g'$.}
3609: %
3610: relations
3611: \begin{gather}
3612: ds^2|_{N_F} \sim dt_F^2 + p_F^* ds^2_F
3613: \label{eqnMetricalCollar} \\
3614: \intertext{and}
3615: h|_{N_F} \sim p_F^* (h|_F) \notag
3616: \end{gather}
3617: uniformly for all codimension $1$ closed boundary faces $F$.  \textup(This
3618: holds in particular if $Y$ is compact.\textup)  Then the natural map induces
3619: an isomorphism
3620: \begin{equation*}
3621: H_{(2),b}(Y;\EE) \longtildearrow H_{(2)}(Y;\EE).
3622: \end{equation*}
3623: \end{lem*}
3624: 
3625: \begin{proof}
3626: Let $\eta(t)$ be a smooth function on $[0,1]$ which is identically 1 on
3627: $[0,1/4]$ and has support in $[0,1/2)$.  Assume first that $Y$ has a smooth
3628: boundary $F$ and set $t=t_F$ and $p=p_F$.
3629: As in \cite{refnCheeger}, the homotopy operator
3630: \begin{equation*}
3631: H\o = 2 \eta(t) \int_0^{1/2}\!\left(\int_a^{t} \iota_{\partial/\partial
3632: t}\o\right) da 
3633: \end{equation*}
3634: yields a homotopy formula 
3635: \begin{equation}
3636: dH + H d = I-P \label{eqnHomotopyFormula}
3637: \end{equation}
3638: on smooth forms, where
3639: \begin{equation*}
3640: P\o = (1-\eta(t))\o + 2 \eta(t)\, p^*\! \int_0^{1/2}\!\!(\o|_{a\times
3641: F})\, da  -  
3642: 2 \eta'(t)\,dt\wedge \int_0^{1/2} \!\left(\int_a^t \iota_{\partial/\partial t}
3643: \o \right) da
3644: \end{equation*}
3645: Given our hypotheses on the metrics, an easy estimate using the
3646: Cauchy-Schwarz inequality demonstrates that $H$ and $P$ are bounded in
3647: the usual $L^2$ norm.  Thus \eqref{eqnHomotopyFormula} holds on $\Dom \dbar$.
3648: On the other hand,
3649: \begin{gather*}
3650: \|H\o|_F\| = 2 \| \int_0^{1/2}\! \left(\int_a^{0} \iota_{\partial/\partial
3651: t}\o\right) da \| \le C \|\o\| \\
3652: \intertext{and}
3653: \|P\o|_F\| =  2 \| \int_0^{1/2}\!\!(\o|_{a\times F})\, da\| \le C
3654: \|\o\|.
3655: \end{gather*}
3656: Thus $P$ defines a bounded map $\Dom\dbar\to \Dom\dbar_b$ and
3657: \eqref{eqnHomotopyFormula} holds on $\Dom\dbar_b$; this proves the lemma
3658: for the case of a single boundary face.  In general one performs this
3659: construction for each codimension 1 closed boundary face and composes the
3660: homotopies and projections in the usual way \cite[\S15]{refndeRham}; the
3661: hypotheses \eqref{eqnNeighborhoodCompatibility} ensure that the
3662: $t_F$-invariance produced by $P_F$ is not disturbed by $P_{F'}$ or
3663: $H_{F'}$.
3664: \end{proof}
3665: 
3666: \begin{rem}
3667: \label{ssectGeneralBoundaryNorm}
3668: If $Y$ is not compact then analogous results may be proven even if
3669: \eqref{eqnMetricalCollar} fails.  For example, if instead of
3670: \eqref{eqnMetricalCollar} we have
3671: \begin{equation*}
3672: ds^2|_{N_F} \sim (p_F^*w_F)^2 dt_F^2 + p_F^* ds^2_F \ ,
3673: \end{equation*}
3674: where $w_F>0$ is a bounded function on $F$, then the lemma continues to
3675: hold provided that $\lvvv\cdot\rvvv$ is defined using an $L^2$-norm for $F$
3676: which is weighted by $w_F^2$  (and a product weight for faces of higher
3677: codimension).  Even weight functions not independent of $t_F$ may be
3678: handled---for a nontrivial example and application see
3679: \cite[\S\S2--4]{refnSaperIsol}. (In particular Lemma ~3.2 of
3680: \cite{refnSaperIsol} is the analogue of the lemma above.)
3681: \end{rem}
3682: 
3683: \subsection{}
3684: \label{ssectLtwoBoundarySpace}
3685: Since $\dbar_b$ is not closed, we need to be careful about the meaning of
3686: the adjoint operator $\dbar_b^*$ before extending the spectral vanishing
3687: criterion of Proposition ~\ref{ssectSpectralVanishingLtwoCohomology} to
3688: $H_{(2),b}^i(Y;\EE)$.  Let $L_{2,b}(Y;\EE)$ denote the Hilbert space
3689: completion of $A_c(Y;\EE)$ with respect to $\lvvv\cdot\rvvv$; by
3690: associating to a form $\o\in A_c(Y;\EE)$ the collection $(\o|_F)_F$ of
3691: restrictions to faces, we obtain an isometry $L_{2,b}(Y;\EE) \cong
3692: \bigoplus_F L_2(F;\EE|_F)$.  In this way $\dom \dbar_b$ may be viewed as a
3693: subset of $L_{2,b}(Y;\EE)$ and the map $(\o|_F)_F\mapsto (d\o|_F)_F$ is a
3694: closed unbounded operator.  Define $\dbar_b^*$ to be the adjoint of this
3695: operator in the sense of unbounded operators on $L_{2,b}(Y;\EE)$.
3696: 
3697: To understand $\dbar_b^*$ better, assume that $\psi=(\psi_F)_F\in \dom
3698: \dbar_b^*$ with each $\psi_F$ a smooth form on $F$. Then from the relation
3699: $(\dbar_b^*\psi,\o)_b= (\psi,\dbar_b \o)_b$ which is valid for all
3700: $\o\in\dom \dbar_b$, but in particular for all $\o\in
3701: A_c(Y;\EE)$, it is easy to see that
3702: \begin{equation*}
3703: (\dbar_b^* \psi)_F = \CoDifferential_F\psi_F + \sum_{ \Ftilde \supset F}
3704: (\iota_{\v_F^{\Ftilde}} \psi_{\Ftilde})|_F.
3705: \end{equation*}
3706: Here $\CoDifferential_F$ is the formal adjoint to $d$ on $F$ and the sum is
3707: over boundary faces $\Ftilde$ for which $F$ is a codimension $1$ face with
3708: unit normal vector $\v_F^{\Ftilde}$.
3709: 
3710: Note that $\psi$ belonging to $\dom \dbar_b^*$ does not necessarily imply
3711: that the various $\psi_F$ satisfy Neumann boundary conditions.  Furthermore
3712: even if $\psi$ is induced from a smooth form on $Y$, the same is not
3713: necessarily true for $\dbar_b^*\psi$.
3714: 
3715: \subsection{Spectral vanishing criterion for $H_{(2),b}(Y;\EE)$}
3716: \label{ssectSpectralVanishingBoundaryValues}
3717: To state the vanishing criterion, define
3718: \begin{multline*}
3719: \H_{\epsilon,b}(Y;\EE)= \{\,\o\in \dom \dbar_b \cap\dom
3720: \dbar_b^*,\ \supp \o \text{ compact} \mid  \\
3721: \lvvv \dbar_b \o \rvvv^2 + \lvvv
3722: \dbar_b^*\o\rvvv^2 \le \epsilon\lvvv \o \rvvv^2 \,\}.
3723: \end{multline*}
3724: Note that from now on we will resume viewing $\dom \dbar_b$ (and hence $\o$
3725: above) as being contained in $L_{2}(Y;\EE)$.  Even so, $\dbar_b^*\o$
3726: must still be viewed as an element of $L_{2,b}(Y;\EE)$ as in the
3727: previous paragraph.  We give $\ker \dbar_b$ and $\range \dbar_b$ the
3728: topology coming from the $\lvvv\cdot\rvvv$-norm and let
3729: $H_{(2),b}(Y;\EE)$ have the quotient topology.
3730: 
3731: \begin{prop*}
3732: Assume that $Y$ is complete as a metric space.  Then
3733: $\H_{\epsilon,b}^i(Y;\EE) = 0$ for some $\epsilon>0$ if and only if
3734: $H_{(2),b}^i(Y;\EE)=0$ and $H_{(2),b}^{i+1}(Y;\EE)=0$ is Hausdorff.
3735: \end{prop*}
3736: 
3737: \begin{proof}
3738: The desired equivalence would follow from Banach's closed range theorem
3739: \cite[\S1.1]{refnHormander} if it were not for the restriction that $\supp
3740: \o$ be compact in the definition of $\H_{\epsilon,b}^i(Y;\EE)$.  To see
3741: that this doesn't matter, first note that by Cohn-Vossen's generalization
3742: \cite[2.4]{refnBallmann} of the Hopf-Rinow theorem, the completeness of $Y$
3743: implies that the function $r(x)$ (distance in $Y$ from a fixed point) has
3744: compact sublevel sets.  Then (following \cite{refnGaffney} except we don't
3745: bother regularizing) use $r(x)$ to create a sequence of compactly supported
3746: Lipschitz functions $\eta_k$ such that $0\le \eta_k \le 1$, $|d\eta_k|\le
3747: C$, and every compact $K\subseteq Y$ is contained in $\eta_k^{-1}(1)$ for
3748: $k$ sufficiently large.  For $\o\in \dom \dbar_b \cap\dom \dbar_b^*$, it
3749: follows that $\eta_k\o\to \o$, $\dbar_b(\eta_k\o)\to \dbar_b\o$, and
3750: $\dbar_b^*(\eta_k\o)\to \dbar_b^*\o$ in the $\lvvv\cdot\rvvv$ norm.  Thus
3751: the existence of a nonzero $\o\in \dom \dbar_b \cap\dom \dbar_b^*$
3752: satisfying $\lvvv \dbar_b \o \rvvv^2 + \lvvv \dbar_b^*\o\rvvv^2 \le
3753: \epsilon\lvvv \o \rvvv^2$ would imply that $\H_{2\epsilon,b}^i(Y;\EE) \neq
3754: 0$.
3755: \end{proof}
3756: 
3757: \begin{rem*}
3758: The proposition would remain true if in the definition of
3759: $\H_{\epsilon,b}(Y;\EE)$ we restrict to $\o$ smooth (as for ordinary
3760: $L^2$-cohomology).  The proof requires a generalization of Friedrich's
3761: regularization techniques to handle boundary values; since we will not need
3762: this stronger result, we will not go into details.
3763: \end{rem*}
3764: 
3765: \subsection{$L^2$-cohomology with boundary values of $\Xhat$}
3766: \label{ssectLtwoBoundaryValuesXhat}
3767: In the next section we will define a Riemannian metric on $\Xbar$ (actually
3768: on a space $\Xbar_t$ diffeomorphic to $\Xbar$) whose restriction to
3769: $Y_P\subseteq \partial \Xbar$ is $N_P$-invariant.  Thus there is an induced
3770: Riemannian metric on every strata $X_P$ of $\Xhat$.  Even though $\Xhat$ is
3771: not a manifold with corners, we can define the operators $\dbar_{\Xhat}$
3772: and $\dbar_{\Xhat,b}$ as well as the $L^2$-cohomology groups
3773: $H_{(2)}(\Xhat;\EE)$ and $H_{(2),b}(\Xhat;\EE)$ just as for $\Xbar$ but
3774: with the following differences:
3775: \begin{itemize}
3776: \item Forms in the domain of $\smash[t]{\dbar}_{\Xhat}$ and
3777: $\smash[t]{\dbar}_{\Xhat,b}$ belong to the complex of special differential
3778: forms $A\sp(\Xhat;\EE)$ as opposed to $A(\Xbar;\EE)$.
3779: \item The $\lvvv\cdot\rvvv$-norm  is defined using the restrictions
3780: $\o\mapsto \o|_{X_P}$ from \eqref{eqnRestrictSpecialForms} as opposed to
3781: $\o\mapsto \o|_{Y_P}$.
3782: \end{itemize}
3783: All the results in the preceding subsections extend to this context (and
3784: likewise if $\Xhat$ is replaced by any domain $U\subset \Xhat$).
3785: 
3786: \section{$L^2$-cohomology of $\L$-modules}
3787: \label{sectLTwoCohomologyLModules}
3788: In this section we will represent the cohomology $H(\Xhat;\M)$ of an
3789: $\L$-module $\M$ by a variant of $L^2$-cohomology.  To do this, we need to
3790: consider $\Xbar_t \subset X$, a diffeomorphic image of $\Xbar$ which was
3791: constructed in \cite{refnSaperTilings}, and replace $\Xhat$ by the
3792: corresponding $\Xhat_t$.  We state a vanishing criterion for this
3793: $L^2$-cohomology analogous to those in Propositions
3794: ~\ref{ssectSpectralVanishingLtwoCohomology} and
3795: \ref{ssectSpectralVanishingBoundaryValues}.
3796: 
3797: We also introduce a variant of $\L$-modules on $\Abar_P\times X_P$ and
3798: their $L^2$-cohomology.
3799: 
3800: \subsection{Locally symmetric metrics}
3801: \label{ssectMetrics}
3802: We will work with a fixed basepoint $x\in D$.  Let $KA_G\subseteq
3803: G(\RR)$ be the stabilizer of the basepoint $x$ and let $\theta$ be the
3804: Cartan involution associated to $K$ (see for example
3805: \cite[V.24.6]{refnBorelLAG} or \cite[1.6]{refnBorelSerre} for the
3806: nonconnected and nonsemisimple case) with Cartan decomposition
3807: $\mathfrak g=\k+\p+\sa_G$.  Fix an $\Ad G(\RR)$-invariant,
3808: $\theta$-invariant nondegenerate bilinear form $B$ on $\mathfrak g$
3809: such that $(X,Y) \mapsto -B(X,\theta Y)$ is a positive definite inner
3810: product on $\mathfrak g$.  (On the derived algebra one may take $B$ to
3811: be the Killing form.)  The restriction of $B$ to $\p+\sa_G$ yields an
3812: $\Ad KA_G$-invariant inner product on $T_x D$ which may be extended to
3813: a $G(\RR)$-invariant metric on $D$; this descends to a locally
3814: symmetric Riemannian metric on $X=\G\back D$.  In addition one obtains
3815: from $B$ left-invariant metrics on all Lie subgroups of $G(\RR)$.
3816: 
3817: If $(E,\sigma)$ is a regular representation of $G$ let $\EE = D \times_{\G}
3818: E$ be the corresponding flat bundle on $X$.  An {\itshape admissible\/}
3819: inner product on $E$ is a positive definite Hermitian bilinear form on $E$
3820: which is $K$-invariant and for which $\p+\sa_G$ acts via self-adjoint
3821: operators; equivalently it is a form for which $\sigma(g)\sigma(\theta
3822: g)^*=\id_E$ for $g\in G(\RR)$.  Admissible metrics always exist (see
3823: \cite[p.~ 375]{refnMatsushimaMurakami} in the case where $G(\RR)$ is
3824: connected and semisimple; the generalization to our situation is clear).
3825: Such an admissible inner product on $E$ induces a metric on $\EE$; in the
3826: case that $E$ is isotypical it is given by the formula $| (gKA_G,v) | =
3827: |\xi_E^k(g)|^{\frac1k} \cdot |\sigma(g^{-1})v|$, where $\xi_E$ is the
3828: character by which $S_G$ acts on $E$ and $k\in\NN$ is such that $\xi_E^k$
3829: extends to a character on $G$.
3830: 
3831: The above construction may be applied to all $L_P$ (using the basepoint
3832: $A_P \geo N_Px\in D_P$, corresponding maximal compact subgroup
3833: $K_P\subseteq L_P(\RR)$ and the bilinear form on $\levi_P$ induced from
3834: $B$) in order to define locally symmetric metrics on the $X_P$ and metrics
3835: on flat bundles $\EE_P$ associated to regular representations $E_P$ of
3836: $L_P$.
3837: 
3838: For a parabolic $P$ there is (analogously to
3839: \S\ref{ssectNilmanifoldFibrations}) a flat nilmanifold fibration $\G_P\back
3840: D \to Z_P \equiv \G_{L_P}\back (N_P(\RR)\back D)$.  The space
3841: $N_P(\RR)\back D$ is naturally a homogeneous space for $L_P(\RR)$ with the
3842: stabilizer of a point being a maximal compact subgroup times $A_G$.  The
3843: metric on $D$ induces a $L_P(\RR)$-invariant metric on $N_P(\RR)\back D$
3844: and hence a metric on $Z_P$.  There is a trivial principal $A_P^G$-bundle
3845: $Z_P \to X_P$ given by geodesic retraction; canonical trivializing sections
3846: are given by quotients of $L_P(\RR)$-orbits.  The metric on $Z_P$
3847: decomposes under a canonical trivialization into the sum of an invariant
3848: metric on $A_P^G$ and the locally symmetric metric on $X_P$.  Given a
3849: regular representations $E_P$ of $L_P$, let $\EEtilde_P = (N_P(\RR)\back D)
3850: \times_{\G_{L_P}} E_P$ be the corresponding flat bundle on $Z_P$;
3851: alternatively $\EEtilde_P$ is the pullback of $\EE_P$ on $X_P$.  The
3852: admissible inner product on $E_P$ induces a metric on $\EEtilde_P$ by the
3853: formula $|(rK_PA_P,v)| = |\sigma(r^{-1})v|$ for all $r \in M_P(\RR)
3854: A_P^G$; note that this is {\itshape not\/} the pullback of the metric on
3855: $\EE_P$ unless $A_P^G$ acts trivially.
3856: 
3857: \subsection{Diffeomorphisms of $\Xbar$ into $X$}
3858: \label{ssectTilings}
3859: By \S\ref{ssectGeodesicRetraction} we may identify $D$ with $A_P^G \times
3860: e_P$ such that the $\lsp0P(\RR)$-orbit of the fixed basepoint $x$ is
3861: identified with $\{1\}\times e_P$.  This induces identifications $\G_P\back
3862: D \cong A_P^G \times Y_P$ and $Z_P \cong A_P^G\times X_P$.
3863: 
3864: By \cite[Theorem~6.1]{refnSaperTilings}%
3865: %
3866: \footnote{In fact \cite{refnSaperTilings} constructs a disjoint
3867: decomposition of $D$ (indexed by parabolic $\QQ$-subgroups) called a
3868: {\itshape tiling\/} of which $\Dbar_t$ is the central tile corresponding to
3869: $G$; see in particular Definition ~2.1, Theorem ~5.7, and Theorem ~6.1.
3870: Note the differences in notation: in \cite{refnSaperTilings}, $X$ refers to
3871: the symmetric space itself rather than its arithmetic quotient.  More
3872: significantly, in \cite{refnSaperTilings} $X_P$ denotes the piece of the
3873: tiling corresponding to $P$, as opposed to a stratum of $\Xhat$ as in the
3874: current paper.}
3875: %
3876: there exists a family (depending on $t\in A^G$ sufficiently dominant) of
3877: $\G$-equivariant piecewise-analytic diffeomorphisms
3878: \begin{equation*}
3879: s_t\colon  \Dbar\longtildearrow \Dbar_t\subseteq D\ ,
3880: \end{equation*}
3881: such that $\Xbar_t = \G\back \Dbar_t\subseteq D$ is a compact subdomain
3882: with corners.  For all%
3883: %
3884: \footnote{Actually equation \eqref{eqnBoundaryLocation} is only valid for
3885: $P$ belonging to a fixed set of representatives of $\G$-conjugacy classes
3886: of parabolic $\QQ$-subgroups; we will assume such representatives have been
3887: chosen.  For other $P$ one has $e_{P,t}\subseteq \{t_Pb_P\}\times e_P$
3888: where $b_P\in A_P^G$ is a certain parameter needed to make the construction
3889: $\G$-invariant; this is because $\gamma\in \G$ may move the fixed basepoint
3890: $x$.  See \cite[\S2]{refnSaperTilings}, particularly \S\S2.6, 2.7.}
3891: %
3892: $P\in\Pl$ each diffeomorphism $s_t$ satisfies:
3893: \begin{equation}
3894: e_P \xrightarrow{s_t|_{e_P}} e_{P,t}\equiv 
3895: s_t(e_P) \text{ is $N_P(\RR)$-equivariant}
3896: \label{eqnNPequivariance}
3897: \end{equation}
3898: and
3899: \begin{equation}
3900: e_{P,t} \subseteq  \{t_P\}\times e_P \label{eqnBoundaryLocation}
3901: \end{equation}
3902: (where $t_P\in A_P^G$ is the image
3903: of $t$ under the canonical projection $A^G\cong A_P^G\times A^P \to A_P^G$).
3904: We obtain an induced diffeomorphism
3905: \begin{equation*}
3906: \Xbar \longtildearrow \Xbar_t = \G\back \Dbar_t \subseteq X\ .
3907: \end{equation*}
3908: 
3909: \subsection{}
3910: We can describe $\Xbar_t\subseteq X$ more precisely by using a cylindrical
3911: cover of $X$.  Given $s_P\in A_P^G$ and $t\in A^G$, set
3912: \begin{align*}
3913: A_P^G(s_P) &= \{\, a\in A_P^G\mid a^\al > s_P^\al \text{ for all
3914: $\al\in\D_P$}\,\}, \text{ and} \\
3915: \langle A_P^G\rangle_t &= \{\, a\in A_P^G \mid a^\b \le t^\b \text{ for all
3916: $\b\in\Dhat_P$}\,\}.
3917: \end{align*}
3918: For $\Omega_P\subseteq Y_P$ open, relatively compact, and
3919: $N_P(\RR)$-invariant, let $W_P = A_P^G(s_P) \times \Omega_P\subseteq
3920: A_P^G\times Y_P \cong \G_P\back D$ be the associated {\itshape cylindrical
3921: set}.  We always assume that $s_P$ is sufficiently dominant (depending on
3922: $\Omega_P$) so that $W_P$ may be identified with its image in $X=\G\back
3923: D$.  For appropriate $s_P$ and $\Omega_P$ we obtain a {\itshape cylindrical
3924: cover\/} $\{W_P\}_{P\in\Pl}$ of $X$.  Given any cylindrical cover
3925: $\{W_P\}$, the intersections $\Xbar_t \cap W_P$ form a cover of $\Xbar_t$;
3926: for $t$ sufficiently dominant one can prove \cite[Theorem
3927: ~5.7]{refnSaperTilings} that $\Xbar_t \cap W_P\neq\emptyset$ and that
3928: \begin{equation}
3929: \Xbar_t \cap W_P = (\langle A_P^G\rangle_t \times Y_P)\cap W_P\ .
3930: \label{eqnAdaptedTiling}
3931: \end{equation}
3932: We set $\Wbar_{P,t} \equiv \Xbar_t \cap W_P$.
3933: 
3934: \subsection{}
3935: \label{ssectMetrizeRetract}
3936: The diffeomorphism $\Xbar \tildearrow \Xbar_t$ induces diffeomorphisms
3937: $Y_P \tildearrow Y_{P,t} = \G_P\back e_{P,t}\subseteq \G_P\back D$ of the
3938: boundary faces.   For each $P$ this diffeomorphism descends by
3939: \eqref{eqnNPequivariance} to
3940: \begin{equation}
3941: X_P \longtildearrow X_{P,t} \subseteq Z_P\ ,
3942: \label{eqnStratumRetract}
3943: \end{equation}
3944: where $X_{P,t}$ is formed from $Y_{P,t}$ by collapsing the
3945: $N_P(\RR)'$-fibers of the nilmanifold fibration.  We set 
3946: \begin{equation*}
3947: \Xhat_t= \coprod_P X_{P,t}
3948: \end{equation*}
3949: and equip it with the topology so that \eqref{eqnStratumRetract} induces a
3950: homeomorphism $\Xhat\tildearrow \Xhat_t$.  The closure of $X_{P,t}$ will be
3951: denoted $\Xhat_{P,t}$.
3952: 
3953: The locally symmetric metric on $Z_P=A_P^G\times X_P$ induces on each stratum
3954: $X_{P,t}$ of $\Xhat_t$ a metric which extends smoothly to its Borel-Serre
3955: compactification $\Xbar_{P,t}$.  Given a regular representation $E_P$ of
3956: $L_P$ there is a natural isomorphism of flat bundles
3957: \begin{equation*}
3958: \EE_P \longtildearrow \EE_{P,t} \equiv \EEtilde_P|_{X_{P,t}}
3959: \end{equation*}
3960: over the diffeomorphism \eqref{eqnStratumRetract}.  As in
3961: \S\ref{ssectMetrics}, the choice of an admissible inner product on $E_P$
3962: induces a metric on $\EEtilde_P$, and hence a metric on $\EE_{P,t}$ that
3963: also extends smoothly over $\Xbar_{P,t}$.
3964: 
3965: \subsection{}
3966: \label{ssectLtwoMethods}
3967: Let $\M=(E_\cdot,f_{\cdot\cdot})$ be an $\L$-module on $\Xhat$.  In
3968: \S\ref{sectRealizationLModules} we constructed an incarnation
3969: $\A_{\Xhat}(\M)$ of the realization using special differential forms; this
3970: was a complex of fine sheaves on $\Xhat$.  The cohomology $H(\Xhat;\M)$ is
3971: the cohomology of the global sections of $\A_{\Xhat}(\M)$.
3972: 
3973: We now consider an alternate incarnation and realization which is a complex
3974: of sheaves on $\Xhat_t$ instead of $\Xhat$.  Namely we set $\A_P(E_P) =
3975: \Asp(\Xhat_{P,t};\EE_{P,t})$ instead of $\Asp(\Xhat_P;\EE_P)$ and let
3976: $\A_{\Xhat_t}$ be the corresponding realization functor provided by (a
3977: generalization of) Theorem ~\ref{ssectRealizationLModules}.  If $U\subseteq
3978: \Xhat_t$ is a domain define $H(U;\M)$ to be the cohomology of global
3979: sections of $\A_{\Xhat_t}(\M)|_U$.  For $U=\Xhat_t$, clearly
3980: \begin{equation*}
3981: H(\Xhat;\M) \cong H(\Xhat_t;\M)\ .
3982: \end{equation*}
3983: 
3984: Explicitly for $U\subseteq \Xhat_t$ the cohomology $H(U;\M)$ is the
3985: cohomology of the complex
3986: \begin{equation}
3987: \label{eqnCohomologyULModuleComplex}
3988: \left\{
3989: \begin{aligned}
3990: A\sp(U;\M) & = \bigoplus_P
3991:  A\sp(U\cap\Xhat_{P,t};\EE_{P,t}),
3992:  \\
3993: d &= \sum_P d_P + \sum_{P \le Q} \dd_{PQ},
3994: \end{aligned}
3995: \right.
3996: \end{equation}
3997: where $d_P$ is the differential on $A\sp(U\cap\Xhat_{P,t};\EE_{P,t})$ and
3998: $\dd_{PQ}$ is defined by
3999: \begin{equation*}
4000: \dd_{PQ}(\o_Q) = A\sp(U\cap\Xhat_{P,t};f_{PQ}\circ
4001: h_{PQ})(\o_Q|_{U\cap\Xhat_{P,t}}).
4002: \end{equation*}
4003: 
4004: For $\o=(\o_Q)_Q\in A\sp(U;\M)$ define the weighted norm
4005: \begin{equation}
4006: \|\o\|_t^2 = \sum_Q \lvvv t^{-\hsr_Q} \o_Q\rvvv[Q]^2
4007: \label{eqnComplexNorm}
4008: \end{equation}
4009: where the norms on the right-hand side are as in
4010: \S\S\ref{ssectLtwoBoundaryValues} and \ref{ssectLtwoBoundaryValuesXhat},
4011: \begin{equation}
4012: \label{eqnComponentNorm}
4013: \lvvv\o_Q\rvvv[Q]^2 = \sum_{P\le Q} \|\o_Q|_{U\cap X_{P,t}}\|_{U\cap
4014: X_{P,t}}^2\ .
4015: \end{equation}
4016: We let $t^{-\hsr}$ denote the family of weights $(t^{-\hsr_Q})_Q$ and let
4017: \begin{equation*}
4018: L_{2}(U;\M,t^{-\hsr}) = 
4019:    \bigoplus_P  L_{2,b}(U\cap\Xhat_{P,t};\EE_{P,t},t^{-\hsr_Q})
4020: \end{equation*}
4021: denote the Hilbert space completions of $A\sp(U;\M)$ and
4022: $A\sp(U\cap\Xhat_{P,t};\EE_{P,t})$ under the norms
4023: \eqref{eqnComplexNorm} and \eqref{eqnComponentNorm}.
4024: 
4025: For each $P$ the closed unbounded operator $\dbar_{U\cap
4026: \Xhat_{P,t},b,t^{-\hsr_P}}$ from \S\S\ref{ssectLtwoBoundaryValues} and
4027: \ref{ssectLtwoBoundaryValuesXhat} is defined on
4028: $L_{2,b}(U\cap\Xhat_{P,t};\EE_{P,t},t^{-\hsr_Q})$ and has cohomology
4029: $H_{(2),b}(U\cap\Xhat_{P,t};\EE_{P,t},t^{-\hsr_Q})$.  On the other
4030: hand, let $d_{U;\M,t^{-\hsr}}$ be the unbounded operator on
4031: $L_{2}(U;\M,t^{-\hsr})$ corresponding to the subcomplex of
4032: \eqref{eqnCohomologyULModuleComplex} consisting of $\o$ such that $\o$,
4033: $d\o\in L_{2}(U;\M,t^{-\hsr})$ and denote its graph norm closure by
4034: $\dbar= \dbar_{\M}= \dbar_{U;\M,t^{-\hsr}}$.  The cohomology of $\dom
4035: \dbar_{U;\M,t^{-\hsr}}$ is by definition the {\itshape $L^2$-cohomology of
4036: $\M$ over $U$ with weights $t^{-\hsr}$\/} and is denoted
4037: $H_{(2)}(U;\M,t^{-\hsr})$.  In order to relate these
4038: $L^2$-cohomology groups we need the
4039: 
4040: \begin{lem}
4041: \label{ssectDDBoundIndependentT}
4042: The operators $\dd_{PQ}$ are bounded with respect to \eqref{eqnComplexNorm}
4043: independently of $t$.
4044: \end{lem}
4045: \begin{proof}
4046: There are bounds independent of $t$ for both the restriction map and
4047: $f_{PQ}$: the first since $\lvvv\cdot\rvvv[Q]$ incorporates the $L^2$-norm
4048: of boundary values, and the second since it is induced by a map of
4049: $L_P$-modules with admissible inner product.  As for $h_{PQ}$, recall from
4050: the proof of Lemma~\ref{ssectNilmanifoldCohomologyBundle} that this acts on
4051: the coefficients via the composition
4052: \begin{equation*}
4053: \AA\inv(\mathscr N_P^Q(\RR)';\EE_{Q,t}) \xrightarrow{h_x}
4054: \CC(\n_P^Q;E_Q)_x \longrightarrow \HH(\n_P^Q;E_Q)\ .
4055: \end{equation*}
4056: The second map here is induced from a map of $L_P$-modules so it remains to
4057: consider $h_x$ at a point $z\in U\cap X_{P,t}$.  Let $Y_{P,t}^Q$ denote the
4058: Borel-Serre boundary stratum of $\Xbar_{Q,t}$ associated to $P$ and let
4059: $\nil\colon Y_{P,t}^Q \to X_{P,t}$ denote the nilmanifold fibration with
4060: typical fiber $\mathscr N_P^Q(\RR)'$.  The map $h_x$ is induced by the
4061: $\G_{L_P}$-morphism which transfers an invariant form on $\mathscr
4062: N_P^Q(\RR)'$ to one on $N_P^Q(\RR)'$ via $n_x'$ (see
4063: \S\ref{ssectNilmanifoldFibrations}) and then takes its value at the
4064: identity.  If $\psi\in \AA\inv(\mathscr N_P^Q(\RR)';\EE_{Q,t})_z$ is an
4065: invariant form on $\nil^{-1}(z)$, then the pointwise norm $|\psi_y|$ is
4066: independent of $y\in \nil^{-1}(z)$ and thus
4067: \begin{equation*}
4068: |\psi|^2 = \int_{\nil^{-1}(z)} |\psi_y|^2 dy = \Vol(\nil^{-1}(z))|\psi_{y_0}|^2
4069:  = \Vol(\nil^{-1}(z)) |h_x(\psi)|^2.
4070: \end{equation*}
4071: This implies $| h_x |^2 = \Vol(\nil^{-1}(z))^{-1} =
4072: t^{2\hsr_P^Q}\Vol(N_P^Q(\RR)')^{-1}$ and hence $\|h_{PQ}\|_t^2$ is bounded
4073: independently of $t$.
4074: \end{proof}
4075: 
4076: \begin{cor}
4077: There is a decomposition \textup(as vector spaces, not as complexes\textup)
4078: \begin{equation*}
4079: \dom \dbar_{U;\M,t^{-\hsr}}  =
4080: \bigoplus_P \dom \dbar_{U\cap \Xhat_{P,t},b,t^{-\hsr_P}}
4081: \end{equation*}
4082: and a strict operator equality 
4083: \begin{equation*}
4084: \dbar_{U;\M,t^{-\hsr}} = \sum_P
4085: \dbar_{U\cap \Xhat_{P,t},b,t^{-\hsr_P}} + \sum_{P \le Q} \dd_{PQ}\ .
4086: \end{equation*}
4087: \end{cor}
4088: 
4089: For $U=\Xhat$ we have the
4090: 
4091: \begin{cor}
4092: \label{ssectCohomologyLModuleIsLtwoCohomology}
4093: $H(\Xhat_t;\M)\cong H_{(2)}(\Xhat_t;\M,t^{-\hsr})$.
4094: \end{cor}
4095: \begin{proof}
4096: Filter both \eqref{eqnCohomologyULModuleComplex} and $\dom
4097: \dbar_{U;\M,t^{-\hsr}}$ by $\#\D_P\le p$ and compare the resulting
4098: spectral sequences; they are isomorphic at $E_1$ by
4099: \eqref{eqnLtwocohomologyIsOrdinaryForCompact} and Lemma
4100: ~\ref{ssectBoundaryAveraging}.
4101: \end{proof}
4102: 
4103: \subsection{}
4104: \label{ssectSpectralVanishingLModules}
4105: Define
4106: \begin{multline*}
4107: \H_{\epsilon}^i(U;\M,t^{-\hsr})=
4108: \{\,\o\in
4109: \dom \dbar_{U;\M,t^{-\hsr}} \cap\dom \dbar_{U;\M,t^{-\hsr}}^*,\ \supp \o
4110: \text{ compact}
4111: \mid \\
4112: \| \dbar \o \|_t^2 + \| \dbar^*\o\|_t^2 \leq
4113: \epsilon\| \o \|_t^2 \,\}\ .
4114: \end{multline*}
4115: For $U=\Xhat$ we have as in Proposition
4116: ~\ref{ssectSpectralVanishingBoundaryValues}
4117: \begin{prop*}
4118: $\H_{\epsilon}^i(\Xhat;\M,t^{-\hsr})=0$ for some $\epsilon>0$ if  and
4119: only if $H^i_{(2)}(\Xhat;\M,t^{-\hsr})=0$ and
4120: $H^{i+1}_{(2)}(\Xhat;\M,t^{-\hsr})$ is Hausdorff.
4121: \end{prop*}
4122: 
4123: Actually $H_{(2)}(\Xhat;\M,t^{-\hsr})$ is finite-dimensional so the
4124: Hausdorff condition above is vacuous.  However the proposition also applies
4125: in an different context which we will now introduce and where
4126: finite-dimensionality is not assured.
4127: 
4128: \subsection{$\L$-modules on $\langle A_P^G\rangle_t\times X_P$}
4129: \label{ssectLModulesAPTimesXP}
4130: The space $\langle A_P^G\rangle_t \times X_P \subseteq Z_P$ is a manifold
4131: with corners with strata indexed by the parabolic $\QQ$-subgroups $Q\ge P$.
4132: Explicitly the $Q$-stratum is $\langle A_P^G\rangle_t^Q \times X_P$ where
4133: \begin{equation*}
4134: \langle A_P^G\rangle_t^Q = \{\, a\in A_P^G \mid a^\b \le t^\b \text{ for all
4135: $\b\in\Dhat_P\setminus \Dhat_Q$, }a^\b = t^\b \text{ for all
4136: $\b\in\Dhat_Q$}\,\}\ .
4137: \end{equation*}
4138: We may extend the theory of $\L$-modules and their realizations to such
4139: spaces.  In fact the theory becomes much simpler---an $\L$-module
4140: $\M'=(E'_\cdot,f'_{\cdot\cdot})$ on $\langle A_P^G\rangle_t \times X_P$
4141: consists of a family $(E'_Q)_{Q\ge P}$ of graded regular representations of
4142: the {\itshape same\/} group $L_P$ and degree 1 morphisms $f'_{QR}\colon
4143: E'_R \xrightarrow{[1]} E'_Q$.  (Since the space has homotopically trivial
4144: links the functor $H(\n_R^Q;\cdot)$ is replaced throughout by the identity
4145: functor.)  Furthermore special differential forms are no longer needed for
4146: the realization.
4147: 
4148: The strata and the locally constant sheaves associated to $E'_Q$ are
4149: metrized as in \S\ref{ssectMetrizeRetract} by restricting the locally
4150: symmetric metrics on $Z_P$ and $\EEtilde'_Q$.  We will be considering the
4151: $L^2$-space $L_{2}(\langle A_P^G\rangle_t \times X_P; \M')$ (without
4152: weights) and the corresponding $L^2$-cohomology.  Since $\langle
4153: A_P^G\rangle_t \times X_P$ is a complete metric space the analogue of
4154: Proposition ~\ref{ssectSpectralVanishingLModules} holds in this context.
4155: 
4156: \section{Analytic Lemmas}
4157: \label{sectAnalyticLemmas}
4158: The goal of this section is Proposition
4159: ~\ref{ssectParabolicAnalyticVanishingCriterion}, a vanishing criterion for
4160: $H^i(\Xhat;\M)$ in terms of certain $L^2$-cohomology groups of $\langle
4161: A_P^G\rangle_1 \times X_P$ for all $P\in\Pl$.  In view of Proposition
4162: ~\ref{ssectSpectralVanishingLModules} this is accomplished by presenting a
4163: sequence of lemmas that reduce the vanishing of
4164: $\H_{\epsilon}^i(\Xhat;\M,t^{-\hsr})$.  As we proceed, the parameter $t$ will
4165: have to be made successively more dominant.
4166: 
4167: \subsection{}
4168: \label{ssectSpectralMayerVietoris}
4169: Given a cylindrical cover $\{W_{P}\}$ assume $t$ is sufficiently dominant so
4170: that \eqref{eqnAdaptedTiling} holds.  Each boundary stratum
4171: $Y_{Q,t}\cap\Wbar_{P,t}$ of $\Wbar_{P,t}$ for $Q\ge P$ is part of a
4172: nilmanifold fibration with fibers $N_Q(\RR)'$; let $\What_{P,t}\subseteq
4173: \Xhat_t$ denote the result of collapsing these fibers.  We have
4174: the following spectral analogue of the Mayer-Vietoris sequence:
4175: 
4176: \begin{lem*}
4177: Given $\epsilon>0$ there
4178: exists $\epsilon'>0$ and a cylindrical cover $\{W_{P}\}$ of $X$ such
4179: that for all $t$ sufficiently dominant,
4180: $\H_{\epsilon}^i(\What_{P,t};\M,t^{-\hsr})=0$ for all $P$
4181: implies $\H_{\epsilon'}^i(\Xhat_t;\M,t^{-\hsr})=0$.
4182: \end{lem*}
4183: \begin{proof}
4184: For every $c>0$ (to be chosen below) there exists
4185: \cite[Prop.~ 2.1]{refnSaperSternTwo} a
4186: cylindrical cover $\{W_{P}\}$ possessing a partition of unity
4187: $\{\eta_P\}$ such that $|d\eta_P| < c$.  (It is easy to eliminate the
4188: neatness hypothesis.)  Examination of the construction shows that for $P\le
4189: Q$ the function $\eta_P$ is $N_Q(\RR)$-invariant on $W_P\setminus
4190: \bigcup_{P'\nleq Q} W_{P'}$.  But since $W_P\cap Y_{Q,t}$ is empty unless
4191: for $P\le Q$ by
4192: \eqref{eqnBoundaryLocation} and \eqref{eqnAdaptedTiling}, this implies that
4193: $\eta_P|_{Y_{Q,t}}$ is $N_Q(\RR)$-invariant for {\itshape any\/} $P$ and $Q$.
4194: Thus we may pushdown to obtain a
4195: partition of unity $\{\etahat_P\}$ for the covering $\{\What_{P,t}\}$ of
4196: $\Xhat_t$.  Now for $\o\in \H_{\epsilon'}^i(\Xhat_t;\M,t^{-\hsr})$ (with
4197: $\epsilon'>0$ to be chosen below) one may estimate (compare
4198: \cite[Prop.~ 1.4]{refnSaperSternTwo}):
4199: \begin{align*}
4200: \| \dbar(\etahat_P\o) \|_t^2 + \|
4201: \dbar^*(\etahat_P\o)\|_t^2  &\le
4202:  2 \bigl( \| \etahat_P \dbar\o \|_t^2 + \| \etahat_P
4203:  \dbar^*\o\|_t^2 + \| \dbar\etahat_P\wedge \o \|_t^2 + \|
4204:  \dbar\etahat_P\intprod  \o\|_t^2 \bigr) \\
4205: &\le 2 \bigl( \| \dbar\o \|_t^2 + \| \dbar^*\o\|_t^2 + 2c^2
4206:  \| \o \|_t^2\bigr) \\
4207: &< 2(\epsilon' +2c^2)   \| \o \|_t^2 \\
4208: &\le  2(\#\Pl)^2 (\epsilon' +2c^2)  \max_R \|  \etahat_R\o \|_t^2 \le
4209:  \epsilon \max_R \|  \etahat_R\o \|_t^2,
4210: \end{align*}
4211: where for the last line we choose $c>0$ and $\epsilon'>0$ small.  Thus if
4212: $P$ realizes the maximum on the last line we have $\etahat_P\o\in
4213: \H_{\epsilon}^i(\What_{P,t};\M,t^{-\hsr})$.  Since $\o\neq 0$ implies
4214: $\etahat_P\o\neq0$ for this $P$, the lemma is proved.
4215: \end{proof}
4216: 
4217: \subsection{}
4218: \label{ssectSpectralAverging}
4219: Now assume that $W_{P}$ is a cylindrical set for a fixed $P$ and that $t$
4220: is sufficiently dominant so that \eqref{eqnAdaptedTiling} holds.  Let $O_P$
4221: be the image of $\Omega_P$ under the nilmanifold fibration $Y_P\to X_P$.
4222: Each boundary stratum $X_{Q,t}\cap\What_{P,t}$ of $\What_{P,t}$ for $Q\ge
4223: P$ is part of a nilmanifold fibration with fibers $N_P^Q(\RR)'$; let
4224: \begin{equation*}
4225: \What_{P,t} \xrightarrow{\nil} (\langle
4226: A_P^G\rangle_t \cap A_P^G(s_P)) \times O_P 
4227: \subseteq \langle A_P^G\rangle_t \times X_P
4228: \end{equation*}
4229: be the projection obtained by collapsing these fibers.
4230: 
4231: Define a pushforward $\L$-module $\nil_*\M=(E'_\cdot,f'_{\cdot\cdot})$ on
4232: $\langle A_P^G\rangle_t \times X_P\subseteq Z_P$ in the sense of
4233: \S\ref{ssectLModulesAPTimesXP} by $E'_Q = H(\n_P^Q;E_Q)$ and $f'_{QR}=
4234: H(\n_P^Q;f_{QR})$ and set $\nil_*\M\otimes \CC_{\hsr_P} = (E'_\cdot\otimes
4235: \CC_{\hsr_P},f'_{\cdot\cdot}\otimes \id_{\CC_{\hsr_P}})$.
4236: 
4237: \begin{lem*}
4238: Given $\epsilon>0$ and a cylindrical set $W_P$
4239: there exists $\epsilon'>0$ and an injective map
4240: \begin{equation*}
4241: \H_{\epsilon'}^i(\What_{P,t};\M,t^{-\hsr}) \longrightarrow
4242: \H_{\epsilon}^i( \langle A_P^G\rangle_t \times X_P; \nil_*\M\otimes\CC_{\hsr_P})
4243: \end{equation*}
4244: for all $t$ sufficiently dominant.
4245: \end{lem*}
4246: \begin{proof}
4247: For $Q\ge P$ let $\nil_{Q*}$ be the operator on
4248: $L_{2}(X_{Q,t}\cap \What_{P,t};\EE_{Q,t},t^{-\hsr_Q})$ which orthogonally
4249: projects onto $N_P^Q(\RR)$-invariant and $\n_P^Q$-harmonic forms.
4250: Zucker \cite[(4.24)]{refnZuckerWarped} shows there
4251: is a homotopy formula 
4252: \begin{equation}
4253: \dbar_Q H_Q + H_Q \dbar_Q = I - \nil_{Q*} \label{eqnQHomotopy}
4254: \end{equation}
4255: with $H_Q$ and $\nil_{Q*}$ bounded in the $L^2$ norm; both $\|H_Q\|_Q$ and
4256: $\|\nil_{Q*}\|_Q$ are independent of $t$ (though dependent on
4257: $W_P$).  In fact since the
4258: boundary faces of $\Xbar_{Q,t}\cap \What_{P,t}$ contain full
4259: $N_P^Q(\RR)$-fibers, the result holds for the $\lvvv\cdot\rvvv[Q]$ norm
4260: and we obtain a homotopy formula for $\dbar_{Q,b,t^{-\hsr_Q}}$.
4261: 
4262: We need the compatibility conditions
4263: \begin{equation}
4264: \nil_{Q*} \dd_{QR} = \dd_{QR} \nil_{R*} \qquad \text{and}
4265:              \qquad H_Q \dd_{QR} = - \dd_{QR} H_R \label{eqnCompatibility}
4266: \end{equation}
4267: (the sign is due to the fact that $\dd_{QR}$ is degree 1 in the
4268: coefficients).  To see that this can be arranged, note that Zucker's
4269: construction for $H_Q$ depends on the choice of a central series defined
4270: over $\QQ$,
4271: \begin{equation*}
4272: 1 = N_0^Q \subset N_1^Q \subset \dots \subset N_\v^Q = N_P^Q(\RR),
4273: \label{eqnCentralSeries}
4274: \end{equation*}
4275: and unit $\QQ$-root vectors $Z_j^Q\in \n_j^Q$ which satisfy $\n_j^Q =
4276: \n_{j-1}^Q + \RR\cdot Z_j^Q$.
4277: Choose such a central series and root vectors for $N_P(\RR)$ (the case
4278: $Q=G$) and use the induced data for each $N_P^Q(\RR)$.  (Some of the
4279: vectors $Z_j^Q$ will now be zero but those stages in the series may be
4280: omitted.)  One may now verify from the description in
4281: \cite[\S4(c)]{refnZuckerWarped} that \eqref{eqnCompatibility} holds.
4282: 
4283: By summing \eqref{eqnQHomotopy} over $Q\ge P$ we obtain
4284: \begin{equation}
4285: \dbar_{\M}H + H \dbar_{\M} = I - \nil_*, \label{eqnFullHomotopy}
4286: \end{equation}
4287: where $H=\sum_{Q\ge P} H_Q$, $\nil_* = \sum_{Q\ge P} \nil_{Q*}$, and the
4288: $\dd_{QR}$ terms vanish due to \eqref{eqnCompatibility}.
4289: Now for $\o\in \H_{\epsilon'}^i(\What_{P,t};\M,t^{-\hsr})$ calculate:
4290: \begin{align*}
4291: \| \dbar\o \|_t^2 + \| \dbar^*\o\|_t^2 &\le \epsilon'\| \o
4292: \|_t^2 \\
4293: &\le \epsilon' ( \|  \nil_*\o \|_t^2 + 
4294:                  \| (I-\nil_*) \o \|_t^2 ) \\
4295: &\le \epsilon'  \| \nil_*\o \|_t^2 + 
4296:      2\epsilon'\| H\|_t^2\cdot (\| \dbar\o \|_t^2 +
4297:                               \| \dbar^*\o\|_t^2 );
4298: \end{align*}
4299: the final line uses \eqref{eqnFullHomotopy}---see  \cite[Prop.~ 1.6]{refnSaperSternTwo}.  This estimate shows that $\nil_*$ is injective on
4300: $\H_{\epsilon'}^i$ provided we choose $\epsilon'>0$ so that (say)
4301: $2\epsilon'\| H\|_t^2 \le1/2$.  Furthermore in this case the last term
4302: may be absorbed into the left-hand side and we obtain
4303: \begin{equation*}
4304: \| \dbar \nil_*\o \|_t^2 + \| \dbar^*\nil_*\o\|_t^2 \le
4305: \| \nil_*\|_t^2 \cdot(\| \dbar\o \|_t^2 + \| \dbar^*\o\|_t^2) \le
4306: 2\epsilon'\| \nil_*\|_t^2 \cdot\| \nil_*\o \|_t^2.
4307: \end{equation*}
4308: This shows that $\nil_*$ takes $\H_{\epsilon'}^i$ into
4309: $\H_{\epsilon}^i$ provided we choose $\epsilon'>0$ so that
4310: $2\epsilon'\| \nil_*\|_t^2\le \epsilon$.
4311: Since $\| H\|_t$ and $\| \nil_*\|_t $ are bounded
4312: uniformly in $t$, $\epsilon'>0$ may be chosen independently
4313: of $t$.
4314: 
4315: To complete the proof, we embed $\nil_*(\Dom \dbar_{\M,t^{-\hsr}})$ into
4316: $\Dom \dbar_{\nil_*\M\otimes \CC_{\hsr_P}}$ by extending an invariant
4317: compactly supported form by zero and applying Lemma
4318: ~\ref{ssectNilmanifoldCohomologyBundle} to the coefficients; we also need
4319: to choose a flat trivializing section of the bundle associated to
4320: $\CC_{\hsr_P}$.  To see that this is an isometry, note that the $L^2$ norm
4321: integrand for the $Q$ summand of $\nil_*(\Dom \dbar_{\M,t^{-\hsr}})$
4322: involves $t^{-2\hsr_Q}$ whereas for $\Dom \dbar_{\nil_*\M\otimes
4323: \CC_{\hsr_P}}$ the coefficient system $\CC_{\hsr_P}$ contributes
4324: $a^{-2\hsr_P}=a^{-2\hsr_Q}\cdot a^{-2\hsr_P^Q}=t^{-2\hsr_Q}\cdot
4325: a^{-2\hsr_P^Q}$. The discrepancy is canceled by the map $\AA\inv(\mathscr
4326: N_P^Q(\RR)';\EE_{Q,t}) \xrightarrow{h_x} \CC(\n_P^Q;E_Q)_x$ which
4327: multiplies norm squared by $a^{2\hsr_P^Q}$ (see the proof of
4328: Lemma~\ref{ssectDDBoundIndependentT}).
4329: \end{proof}
4330: 
4331: \subsection{}
4332: \label{ssectLPIsotypicalDecomposition}
4333: The isotypical decomposition of a regular $L_P$-module induces a
4334: decomposition of $\L$-modules
4335: \begin{equation*}
4336: \nil_*\M = \bigoplus_V (\nil_*\M)_V
4337: \end{equation*}
4338: where $V$ ranges over irreducible regular $L_P$-modules.
4339: \begin{lem*}
4340: For any $\epsilon>0$ and any $t$, the vanishing
4341: $\H_{\epsilon}^i ( \langle A_P^G\rangle_t \times X_P;
4342: \nil_*\M\otimes \CC_{\hsr_P})=0$ is equivalent to the vanishing conditions 
4343: \begin{equation*}
4344: \H_{\epsilon}^i ( \langle A_P^G\rangle_t \times X_P;
4345: (\nil_*\M)_V\otimes \CC_{\hsr_P})=0, \qquad\text{for all $V$.}
4346: \end{equation*}
4347: \end{lem*}
4348: \begin{proof}
4349: If $\o =\sum_V \o_V \in \H_{\epsilon}^i ( \langle A_P^G\rangle_t \times X_P;
4350: \nil_*\M\otimes \CC_{\hsr_P})$ is nonzero, then for at least one $V$, $\o_V\neq
4351: 0$ and $\o_V \in \H_{\epsilon}^i ( \langle A_P^G\rangle_t \times X_P;
4352: (\nil_*\M)_V\otimes\CC_{\hsr_P})$.  For otherwise we could sum the estimates
4353: $\|\dbar \o_V\|^2 + \| \dbar^*\o_V\|^2 >\epsilon\| \o_V \|^2$ and obtain a
4354: contradiction.  The converse is obvious.
4355: \end{proof}
4356: 
4357: \begin{lem}
4358: \label{ssecttIndependence}
4359: The vanishing condition
4360: $\H_{\epsilon}^i ( \langle A_P^G\rangle_t \times X_P;
4361: (\nil_*\M)_V\otimes\CC_{\hsr_P})=0$ is independent of $t$.
4362: \end{lem}
4363: \begin{proof}
4364: Let $\Psi_t\colon  \langle
4365: A_P^G\rangle_1 \times X_P\tildearrow \langle A_P^G\rangle_t \times X_P$ be the
4366: diffeomorphism given by the left action of $t_P$.  Although $\Psi_t$ is an
4367: isometry, the natural isomorphism of coefficient systems
4368: $(\EE'_{Q,1})_V\otimes \CC_{\hsr_P}\tildearrow
4369: \Psi_t^*((\EE'_{Q,t})_V\otimes \CC_{\hsr_P})$ multiplies norm by
4370: $t^{-(\xi_V+\hsr_P)}$,
4371: where $\xi_V$ is the character by which $S_P$ acts on $V$.
4372: Thus the norm of a form depends on
4373: $t$ only via a scaling factor independent of $Q$.  This implies that
4374: $\dbar^* \Psi_t^* = \Psi_t^*\dbar^*$ and that the scaling factor may be
4375: canceled from all terms of the estimate
4376: $\| \dbar\o\|^2 + \| \dbar^*\o\|^2 \le \epsilon \|
4377: \o \|^2$.
4378: \end{proof}
4379: \begin{rem*}
4380: This lemma would not hold if we had not included the weight factors in
4381: \eqref{eqnComplexNorm}.
4382: \end{rem*}
4383: 
4384: \subsection{}
4385: \label{ssectParabolicAnalyticVanishingCriterion}
4386: We finally arrive at the following vanishing criterion for the cohomology
4387: of an $\L$-module.
4388: \begin{prop*}
4389: Let $\M$ be an $\L$-module on $\Xhat$. Assume
4390: \begin{equation}
4391: \begin{gathered}
4392: H_{(2)}^i ( \langle A_P^G\rangle_1 \times X_P;
4393: (\nil_*\M)_V\otimes\CC_{\hsr_P})=0\ ,\text{and}\\
4394: H_{(2)}^{i+1} ( \langle A_P^G\rangle_1 \times X_P;
4395: (\nil_*\M)_V\otimes\CC_{\hsr_P}) \text{ is Hausdorff.}
4396: \end{gathered}
4397: \label{eqnParabolicAnalyticVanishingCriterion}
4398: \end{equation}
4399: for all parabolic $\QQ$-subgroups $P$ and  $V\in \IrrRep(L_P)$.  Then
4400: $H^i(\Xhat;\M)=0$.
4401: \end{prop*}
4402: 
4403: \begin{proof}
4404: By Corollary ~\ref{ssectCohomologyLModuleIsLtwoCohomology} and Proposition
4405: ~\ref{ssectSpectralVanishingLModules} it suffices to show
4406: $\H_{\epsilon}^i(\Xhat;\M,t^{-\hsr})=0$ for some $\epsilon>0$.  For this
4407: vanishing it suffices by Lemmas
4408: ~\ref{ssectSpectralMayerVietoris}--\ref{ssecttIndependence} to show there
4409: exists $\epsilon>0$ such that
4410: \begin{equation*}
4411: \H_{\epsilon}^i ( \langle A_P^G\rangle_1
4412: \times X_P; (\nil_*\M)_V\otimes\CC_{\hsr_P})=0
4413: \end{equation*}
4414: for all $P$ and $V$.  However by Proposition
4415: ~\ref{ssectSpectralVanishingLModules} again (see
4416: \S\ref{ssectLModulesAPTimesXP}) this is implied by
4417: \eqref{eqnParabolicAnalyticVanishingCriterion}.
4418: \end{proof}
4419: 
4420: \section{Vanishing Theorem for $L^2$-cohomology}
4421: \label{sectVanishingTheoremLtwo}
4422: In the course of verifying \eqref{eqnParabolicAnalyticVanishingCriterion}
4423: we will need to apply the following vanishing theorem to the ordinary
4424: global $L^2$-cohomology $H_{(2)}(X_P;\VV)$.  The heart of the theorem
4425: is a calculation originally due to Raghunathan \cite{refnRaghunathan},
4426: \cite{refnRaghunathanCorrection}.  The theorem was proven in
4427: \cite{refnSaperSternTwo} but since it was not stated explicitly in this
4428: form we sketch the proof in this section.
4429: 
4430: \begin{thm}
4431: \label{ssectRaghunathanVanishing}
4432: Let $X = \G\back G(\RR)/KA_G$ be the locally symmetric space associated to
4433: an arithmetic subgroup $\G$ of a reductive algebraic group $G$ defined over
4434: $\QQ$.  Let $E$ be an irreducible regular representation of $G$ and let
4435: $\EE$ be the associated locally constant sheaf on $X$.  Endow $X$ with a
4436: locally symmetric Riemannian metric and let $\EE$ have a metric induced
4437: from an admissible inner product on $E$.
4438: \begin{enumerate}
4439: \item If $(E|_{\lsp0 G})^* \not\cong \overline {E|_{\lsp0 G}}$, then
4440: $H_{(2)}(X;\EE)=0$.
4441: \label{itemRaghunathanVanishingNotTauInvariant}
4442: \item If $(E|_{\lsp0 G})^* \cong \overline {E|_{\lsp0 G}}$, then
4443: $H_{(2)}^i(X;\EE)=0$ for
4444: \begin{equation*}
4445: i \notin [(\dim D - \dim D(E))/2,(\dim D + \dim D(E))/2]\ ,
4446: \end{equation*}
4447: where $\dim D(E)$ is defined in \textup{\S\ref{ssectVcentralizer}}.
4448: Furthermore, $H_{(2)}^i(X;\EE)$ is Hausdorff \textup(and hence
4449: representable by harmonic forms\textup) for $i=(\dim D - \dim
4450: D(E))/2$.
4451: \label{itemRaghunathanVanishingDegree}
4452: \end{enumerate}
4453: \end{thm}
4454: 
4455: \begin{proof}[Sketch of Proof \textup{(compare
4456:     \protect{\cite[\S\S8--10]{refnSaperSternTwo}})}] We will determine a
4457:   sufficient condition on a degree $i$ so that the estimate
4458: \begin{equation}
4459:  \|d\o\|^2 + \|d^*\o\|^2 > \epsilon\|\o\|^2
4460: \label{eqnBasicEstimate}
4461: \end{equation}
4462: holds for all $\o\in A^i_c(X;\EE)$ and some $\epsilon>0$.  Since $X$ is
4463: complete and without boundary, Proposition
4464: ~\ref{ssectSpectralVanishingLtwoCohomology} shows that this estimate for a
4465: given $i$ implies $H_{(2)}^i( X;\EE)=0$ and $H_{(2)}^{i+1}(X;\EE)$ is
4466: Hausdorff.
4467: 
4468: Let $K\subseteq G(\RR)$ be a maximal compact subgroup with associated
4469: Cartan involution $\theta$ and Cartan decomposition $\mathfrak
4470: g=\k+\p+\sa_G$.  Let $\sigma_1$ denote the coadjoint action of $\k_\CC$ on
4471: $\bigwedge \p_\CC^*$ and let $\sigma_2$ denote the action of $\mathfrak
4472: g_\CC$ on $E$.  Since $X\cong \G\back (\lsp 0 G(\RR))^0/K^0$, the complex
4473: $A_c(X;\EE)$ may as usual be isometrically identified with the complex
4474: \cite[I, \S\S1.2, 5.1, VII, \S2.7]{refnBorelWallach}, \cite[
4475: \S4]{refnMatsushimaMurakami}
4476: \begin{equation*}
4477: \left\{
4478: \begin{gathered}
4479: C(\lsp 0\mathfrak g,\k; E \otimes A_c^0( \G\back (\lsp0G(\RR))^0)) \cong
4480: \Hom_\k(\bigwedge \p, E \otimes A_c^0( \G\back (\lsp0G(\RR))^0)) \ , \\
4481: d = \smash[b]{\sum_k} \epsilon(X_k)(\sigma_2(X_k) + X_k)\ ,
4482: \end{gathered}
4483: \right.
4484: \end{equation*}
4485: where $\{X_k\}$ denotes an orthonormal basis of $\p$, acting on $A_c^0(
4486: \G\back (\lsp0G(\RR))^0)$ via differentiation by the corresponding
4487: left-invariant vector field, and $\epsilon(X_k)$ denotes exterior
4488: multiplication by the corresponding element of the dual basis.  Note that
4489: the operator $d$ extends to the larger space $\bigwedge \p_\CC^* \otimes E
4490: \otimes A_c^0( \G\back (\lsp0G(\RR))^0)$ (without the $\k$-invariance
4491: condition) though it is no longer a differential here.
4492: 
4493: Integration by parts and some multi-linear algebra allow one to extract
4494: 0-order terms and estimate that
4495: \begin{equation}
4496: \|d\o\|^2 + \|d^*\o\|^2 \ge (\D_0\o,\o) \ge 0 \label{eqnPestimate}
4497: \end{equation}
4498: \cite[Props.~9.2~ and~9.4]{refnSaperSternTwo} for a certain operator
4499: $\D_0$.  Explicitly
4500: \begin{equation}
4501: \label{eqnAlgebraicLaplacianExpression}
4502: \D_0 = \sigma_2(C) - \frac12\sigma_2(C_\k) + \frac12\sigma_1(C_\k) -
4503: \frac12(\sigma_1\otimes\sigma_2)(C_\k) \ ,
4504: \end{equation}
4505: where $C$ (resp. $C_\k$) denotes the Casimir operator of $\lsp0\mathfrak g$
4506: (resp. $\k$).  In fact $\D_0$ is induced from the algebraic Laplacian on
4507: $C(\lsp0\mathfrak g,\k; E)= \Hom_{\k}(\bigwedge \p, E) \subseteq \bigwedge
4508: \p^*_\CC \otimes E$.
4509: 
4510: We now need some more notation.  Let $\h=\hb_G + \sa_G =
4511: \hb_{G,\k}+\hb_{G,\p} +\sa_G$ be a fundamental $\theta$-stable Cartan
4512: subalgebra of $\mathfrak g$.  Fix a positive system
4513: $\Phi_\k^+=\Phi^+(\k_\CC,\hb_{G,\k\CC})$ of roots for $\k_\CC$ and consider
4514: any $\theta$-stable positive system $\Phi^+=\Phi^+(\mathfrak g_\CC,\h_\CC)$
4515: compatible with $\Phi_\k^+$.  (A specific choice will be made below.)  Let
4516: $\Phi^+(\p_\CC,\hb_{G,\k\CC})$ denote the positive nonzero weights of
4517: $\hb_{G,\k\CC}$ in $\p_\CC$, counted with multiplicity.  (A weight is
4518: ``positive'' here if it is the restriction to $\hb_{G,\k}$ of a positive
4519: root \cite[10.1]{refnSaperSternTwo}.)
4520: %Assume we have a Weyl group
4521: %invariant inner product on the real span of $\Phi^+$.
4522: Clearly
4523: \begin{equation}
4524: \label{eqnDimensionD}
4525: \dim D = 2\cdot\#\Phi^+(\p_\CC,\hb_{G,\k\CC}) + \dim \hb_{G,\p}\ .
4526: \end{equation}
4527: Let $\lambda$ be the highest weight of $E$ and recall the reductive Lie
4528: subgroup $G(\lambda)\subseteq G$ defined in \S\ref{ssectWeightCentralizer}
4529: with positive roots $\Phi^+(\mathfrak g(\lambda)_\CC,\h_\CC)=
4530: \{\,\g\in\Phi^+\mid \g\perp \lambda\,\}$; define
4531: $\Phi^+(\p(\lambda)_\CC,\hb_{G,\k\CC})=
4532: \{\,\al\in\Phi^+(\p_\CC,\hb_{G,\k\CC})\mid \al\perp \lambda\,\}$ similarly.
4533: Note that if $\lambda|_{\hb_{G,\p}}=0$ (so that $G(\lambda)$ is defined
4534: over $\RR$ as in \S\ref{ssectVcentralizer}) then
4535: $\Phi^+(\p(\lambda)_\CC,\hb_{G,\k\CC})$ is indeed the set of positive
4536: $\hb_{G,\k\CC}$-weights of the split part in the Cartan decomposition of
4537: $\mathfrak g(\lambda)$.  In this case
4538: \begin{equation}
4539: \label{eqnDimensionDLambda}
4540: \dim D(\lambda) = 2\cdot\#\Phi^+(\p(\lambda)_\CC,\hb_{G,\k\CC}) + \dim
4541: \hb_{G,\p} \ .
4542: \end{equation}
4543: 
4544: Equation \eqref{eqnAlgebraicLaplacianExpression} shows that for any fixed
4545: irreducible submodule $R$ of $\sigma_1\otimes\sigma_2|_\k$, $\D_0$ acts as
4546: a certain scalar on all forms in $R\otimes A_c^0( \G\back
4547: (\lsp0G(\RR))^0))$.  We need to determine when this scalar can be zero.  It
4548: suffices to assume $R\subseteq S_1\otimes S_2$, where $S_1$ is an
4549: irreducible submodule of $\sigma_1$ with pure degree $i$ and highest weight
4550: $\v_1$, and $S_2$ is an irreducible submodule of $\sigma_2|_\k$ with
4551: highest weight $\v_2$.  From the usual basis of weight vectors for
4552: $\bigwedge^i \p_\CC^*$ we see that $\v_1$ may be expressed as $\sum_{\al\in
4553: A^+} \al - \sum_{\al\in A^-} \al$, where $A^+$, $A^-\subseteq
4554: \Phi^+(\p_\CC,\hb_{G,\k\CC})$ and
4555: \begin{equation}
4556: \#A^+ + \#A^- \le i \le \#A^+ + \#A^- + \dim \hb_{G,\p}\ .
4557: \label{eqnDegreeBounds}
4558: \end{equation}
4559: Choose $\Phi^+$ now so that $\v_2$ (extended by zero on $\hb_{G,\p}$) is
4560: dominant.  A calculation \cite[Prop.~10.2]{refnSaperSternTwo} using the
4561: well-known expression in terms of highest weights for the value of the
4562: Casimir operator acting on an irreducible representation shows that $\D_0$
4563: is zero on $R$ precisely when
4564: \begin{enumerate}
4565: \renewcommand{\theenumi}{\arabic{enumi}}
4566: \renewcommand{\labelenumi}{(\theenumi)}
4567: \item the highest weight of $R$ is $\v_1+\v_2$, \label{itemWeightDecompose}
4568: \item $\lambda|_{\hb_G} = \v_2$ (where $\v_2$ has been extended to be zero
4569: on $\hb_{G,\p}$), and
4570: \label{itemWeightSplitVanish}
4571: \item $(\Phi^+(\p_\CC,\hb_{G,\k\CC})\setminus A^+)$, $A^-\subseteq
4572: \Phi^+(\p(\lambda)_\CC,\hb_{G,\k\CC})$.
4573: \label{itemWeightOrthogonal}
4574: \end{enumerate}
4575: 
4576: If $\D_0$ acts as a nonzero scalar for all $R$ with a fixed $i$ then
4577: \eqref{eqnPestimate} will yield \eqref{eqnBasicEstimate} and so
4578: $H_{(2)}^i(X;\EE)=0$.  So if $H^i_{(2)}(X;\EE)\neq 0$ conditions
4579: \itemref{itemWeightDecompose}--\itemref{itemWeightOrthogonal} must hold for
4580: some $R$.
4581: Condition \itemref{itemWeightSplitVanish} implies $\tau(\lambda|_{\hb_G}) =
4582: \lambda|_{\hb_G}$ (since $\t = \theta$ for a fundamental Cartan subalgebra
4583: and $\theta$-stable positive system) and thus part
4584: \itemref{itemRaghunathanVanishingNotTauInvariant} of the theorem follows
4585: from \S\ref{ssectConjugateSelfContragredient}.  Condition
4586: \itemref{itemWeightOrthogonal} allows one to estimate the possible maximum
4587: and minimums values of $\#A^\pm$; inserting these into
4588: \eqref{eqnDegreeBounds} yields that
4589: \begin{multline}
4590: \#\Phi^+(\p_\CC,\hb_{G,\k\CC}) - \#\Phi^+(\p(\lambda)_\CC,\hb_{G,\k\CC})
4591: \le i \le \\
4592: \#\Phi^+(\p_\CC,\hb_{G,\k\CC})
4593: +\#\Phi^+(\p(\lambda)_\CC,\hb_{G,\k\CC}) + \dim \hb_{G,\p}
4594: \ .\label{eqnCohomologyDegreeBounds}
4595: \end{multline}
4596: This is equivalent to the degree range in part
4597: \itemref{itemRaghunathanVanishingDegree} by applying  \eqref{eqnDimensionD}
4598: and \eqref{eqnDimensionDLambda} and taking the maximum over $\Phi^+$.
4599: \end{proof}
4600: 
4601: \section{Proof of Theorem ~\ref{ssectGlobalVanishing}}
4602: \label{sectProofGlobalVanishing}
4603: 
4604: By Proposition ~\ref{ssectParabolicAnalyticVanishingCriterion} we need to
4605: show for all $P$ and $V$ that
4606: \begin{equation}
4607: H_{(2)}^i ( \langle A_P^G\rangle_1 \times X_P;
4608: (\nil_*\M)_V\otimes\CC_{\hsr_P})=0 \qquad\text{for $i\notin [c(\M),d(\M)]$}
4609: \label{eqnFinalGoal}
4610: \end{equation}
4611: and in addition
4612: \begin{equation}
4613: H_{(2)}^{c(\M)} ( \langle A_P^G\rangle_1 \times X_P;
4614: (\nil_*\M)_V\otimes\CC_{\hsr_P}) \text{ is Hausdorff.}
4615: \label{eqnFinalGoalAdded}
4616: \end{equation}
4617: 
4618: We may write the complex $C$ which  computes
4619: \eqref{eqnFinalGoal} as 
4620: \begin{equation}
4621: \left\{
4622: \begin{aligned}
4623: C &= \bigoplus_{R\ge P} \dom \dbar_{R,b}\otimes
4624: \Hom_{L_P}(V,H(\n_P^R;E_R))\ , \\
4625: d_C &= \sum_{R\ge P} \dbar_{R,b}\otimes
4626: (-1)^{\text{degree}} + \sum_{R' \ge R \ge P} \i_R^{R'*}\otimes
4627: \Hom_{L_P}(V,H(\n_P^{R'};f_{RR'}))\ ,
4628: \end{aligned}
4629: \right.
4630: \label{eqnNPAveragedComplex}
4631: \end{equation}
4632: where $\dbar_{R,b} = \dbar_{\langle A_P^R\rangle_1\times X_P,b,\VV\otimes
4633: \CC_{\hsr_P}}$ and $\i_R^{R'*}\colon \dom \dbar_{R',b} \to \dom
4634: \dbar_{R,b}$ restricts a form to the smaller stratum.
4635: 
4636: Let $P'\ge P$ be the parabolic $\QQ$-subgroup with type $\D_P^{P'}=
4637: \{\,\al\in \D_P \mid \langle\xi_V+\hsr,\al\spcheck\rangle = 0\,\}$ and set
4638: $S=(P,P')$.  Then any $R$ with $P\le R$ belongs to $[\tilde P, \tilde S]$
4639: for a unique $\tilde P\in [P, P']$ and $\tilde S =\tilde P\vee S$:
4640: \begin{equation*}
4641: \vcenter{\xymatrix @ur @M=1pt @R=1.5pc @C=1pc {
4642: {P'} \ar@{-}[rr] \ar@{-}[d] & {} & {G} \ar@{-}[d] \\
4643: {\tilde P} \ar@{-}[r] \ar@{-}[d] & {R} \ar@{-}[r] &
4644: {\tilde S} \ar@{-}[d] \\
4645: {P\rlap{\qquad\qquad .}} \ar@{-}[rr] & {} & {S}
4646: }}
4647: \end{equation*}
4648: (Compare \eqref{eqnParallelogramWithPtilde}.)  For each such $\tilde P$ let
4649: $C(\tilde P)$ be the subquotient complex obtained by restricting the sum in
4650: \eqref{eqnNPAveragedComplex} to be over those $R\in [\tilde P,
4651: \tilde S]$.  It suffices to prove the vanishing and Hausdorff assertions
4652: of \eqref{eqnFinalGoal} and \eqref{eqnFinalGoalAdded} for each complex
4653: $C(\tilde P)$.  (Compare the Fary spectral sequence from Lemma
4654: ~\ref{ssectRelativeLocalCohomologySupportsSS}.)
4655: 
4656: Filter the double complex $C(\tilde P)$ by the degree of the second factor
4657: in \eqref{eqnNPAveragedComplex}; the associated spectral sequence has
4658: \begin{equation}
4659: E_1^{i-k,k} = \bigoplus_{ R\in [\tilde P,\tilde S] }
4660: H^{k}_{(2)}(\langle A_P^R\rangle_1\times
4661: X_P;\VV\otimes\CC_{\hsr_P})\otimes
4662: \Hom_{L_P}^{i-k}(V,H(\n_P^R;E_R)). \label{eqnRawEoneTerm}
4663: \end{equation}
4664: Note that we may use ordinary $L^2$-cohomology here as opposed to
4665: $L^2$-cohomology with boundary values in view of
4666: Lemma~\ref{ssectBoundaryAveraging}.
4667: 
4668: For $\al\in \D_P$, let $A_P^\al$ denote the one-parameter subgroup
4669: corresponding to $\al\spcheck$.  We have a factorization $\langle
4670: A_P^R\rangle_1 = \prod_{\al\in \D_P^R} \langle A_P^{\al}\rangle_1$; for
4671: each $\al$, Zucker's $L^2$-K\"unneth theorem
4672: \cite[(2.34)]{refnZuckerWarped} may be applied to correspondingly factor
4673: out the $L^2$-cohomology of $\langle A_P^{\al}\rangle_1$ in
4674: \eqref{eqnRawEoneTerm} provided it is Hausdorff.  To calculate the result
4675: write $\xi_V+\hsr = \sum_{\al\in\D_P}
4676: \langle\xi_V+\hsr,\al\spcheck\rangle\b_\al$ and recall that
4677: \begin{equation*}
4678: H^{\cdot}_{(2)}(\langle
4679: A_P^\al\rangle_1;\CC_{\langle\xi_V+\hsr,\al\spcheck\rangle\b_\al} )
4680: \cong \begin{cases} 0 & \text{if $\langle\xi_V+\hsr,\al\spcheck\rangle>0$,} \\
4681: 		M[-1] & \text{if $\langle\xi_V+\hsr,\al\spcheck\rangle=0$,} \\
4682: 		\CC   & \text{if $\langle\xi_V+\hsr,\al\spcheck\rangle<0$,}
4683: \end{cases}
4684: \end{equation*}
4685: where $M$ is an infinite dimensional non-Hausdorff space
4686: \cite[Prop. ~3.2]{refnSaperCertain}, \cite
4687: [(2.37)--(2.40)]{refnZuckerWarped}.  Thus let $Q_V\le S$ have type
4688: $\D_P^{Q_V} =\{\,\al\in \D_P \mid \langle\xi_V+\hsr,\al\spcheck\rangle <
4689: 0\,\}$ as in \S\ref{ssectMicroSupport} and set $\tilde Q=\tilde P\vee Q_V$.
4690: We find that
4691: \begin{equation*}
4692: E_1^{i-k,k} = 
4693: H^{k}_{(2)}(\langle A_P^{\tilde P}\rangle_1\times
4694: X_P;\CC_{\xi_V^{-1}}\otimes\VV)\otimes
4695: \biggl(
4696: \bigoplus_{ R\in [\tilde P, \tilde Q] }
4697: \Hom_{L_P}^{i-k}(V,H(\n_P^R;E_R)) \biggr).
4698: \end{equation*}
4699: 
4700: The advantage of this expression is that the differentials in the spectral
4701: sequence act only on the final factor.  These differentials are induced by
4702: $H(\n_P^R;f_{RR'})$ for $\tilde P\le R\lneq R'\le \tilde Q$ and so from the
4703: definitions in \S\ref{ssectFunctorsOnLsheaves} we see that $H^i(C(\tilde
4704: P))$ equals
4705: \begin{equation}
4706: \bigoplus_k
4707: H^{k}_{(2)}(\langle A_P^{\tilde P}\rangle_1\times
4708: X_P;\CC_{\xi_V^{-1}}\otimes\VV)\otimes
4709: \Hom_{L_P}(V, H^{i-k}(\i_P^* \i_{{\tilde P}*}\i_{\tilde P}^* \ihat_{\tilde Q}^!
4710: \M))\ . \label{eqnEinfinity}
4711: \end{equation}
4712: By the $L^2$-K\"unneth theorem this would be isomorphic to
4713: \begin{equation}
4714: \bigoplus_{k_1,k_2} H^{k_1}_{(2)}(\langle A_P^{\tilde P}\rangle_1)\otimes
4715: H^{k_2}_{(2)}(X_P;\VV)\otimes \Hom_{L_P}(V, H^{i-k_1-k_2}(\i_P^* \i_{{\tilde
4716: P}*}\i_{\tilde P}^* \ihat_{\tilde Q}^!  \M))\ ; \label{eqnDecomposedEinfinity}
4717: \end{equation}
4718: if at least one of the first two $L^2$-cohomology factors were
4719: finite-dimensional in all degrees, however we cannot assume this.  Instead
4720: note that by the technique in \cite[(2.29)--(2.34)]{refnZuckerWarped} one
4721: can show that if each factor of \eqref{eqnDecomposedEinfinity} vanishes
4722: outside a certain interval of degrees, then \eqref{eqnEinfinity} vanishes
4723: outside the sum of the intervals.  The first factor is $0$ outside $[0,\dim
4724: \sa_P^{\tilde P}]$, the vanishing interval of the second factor is given by
4725: Theorem ~\ref{ssectRaghunathanVanishing}, and the third factor vanishes
4726: outside $[c_2(V,\tilde P;\M),d_2(V,\tilde P;\M)]$ by definition (see
4727: Proposition ~\ref{ssectAlternateMicroSupport}).  Thus by Proposition
4728: ~\ref{ssectEqualityDegreeRanges} we have vanishing for $i\notin
4729: [c_2(\M),d_2(\M)]= [c(\M),d(\M)]$ which proves \eqref{eqnFinalGoal}.
4730: 
4731: Finally consider \eqref{eqnFinalGoalAdded}.  If $P=\tilde P$ this follows
4732: from the final assertion of Theorem ~\ref{ssectRaghunathanVanishing}.  If
4733: $P \neq \tilde P$ then $H^{k_1}_{(2)}(\langle A_P^{\tilde P}\rangle_1)$
4734: vanishes in degree $0$ (and is non-Hausdorff for $1 \le k_1 \le \dim
4735: \sa_P^{\tilde P}$ \cite[(4.51)]{refnZuckerWarped}) and thus
4736: \eqref{eqnEinfinity} vanishes in degree $c(\M)$ as well.  \qed
4737: 
4738: \specialsection*{Part III. Micro-support Calculations}
4739: 
4740: \section{Micro-support of Weighted Cohomology}
4741: \label{sectMicroSupportWeightedCohomology}
4742: 
4743: \subsection{}
4744: \label{ssectLocalWeightedCohomologyWithAndWithoutProperSupports}
4745: Fix a weight profile $\eta$.  For $V$ an irreducible $L_P$-module, let
4746: $T_V^\eta\ge P$ have type%
4747: %
4748: \footnote{This type corresponds to what in
4749: \cite{refnGoreskyMacPhersonTopologicalTraceFormula} would be denoted
4750: $I_\eta(\xi_V)$.}
4751: %
4752: \begin{equation*}
4753: \{\,\al\in\D_P\mid \langle\xi_V-\eta_P,\b_\al\spcheck\rangle<0\,\}\ .
4754: \end{equation*}
4755: relative to $P$.  Note that $T_V^\eta=P$ if and only if $\xi_V\in \eta_P +
4756: \lsp+\sa_P^{G*}$ and $T_V^\eta =G$ if and only if $\xi_V \in \eta_P
4757: -\Int(\lsp+\sa_P^{G*})$.
4758: 
4759: 
4760: \begin{lem*}
4761: Let $V$ be an irreducible $L_P$-module occurring in $H(\i_P^*\WnC(E))$.
4762: Then $V$ has the form $H^{\l(w)}(\n_P;E)_w$ for some $w\in W_P$.
4763: Furthermore, the natural morphism $\WnC(E) \to \i_{G*}E$ induces
4764: \begin{equation*}
4765: H(\i_P^*\WnC(E))_V \cong \begin{cases}
4766: V[-\l(w)] & \text{if $T_V^\eta = P$,} \\
4767: 0 & \text{otherwise.}
4768: \end{cases}
4769: %\label{eqnLocalWeightedCohomologyMiddleWeightProfile}
4770: \end{equation*}
4771: \end{lem*}
4772: 
4773: \begin{proof}
4774: Apply Proposition ~\ref{ssectLocalWeightedCohomology}.
4775: \end{proof}
4776: 
4777: \begin{prop}
4778: \label{ssectWeightCohomologyLocalCohomologyWithSupports}
4779: Let $V$ be an irreducible $L_P$-module occurring in
4780: $H(\i_P^*\ihat_Q^!\WnC(E))$ for $P\le Q$.  Then $V$ has the form
4781: $H^{\l(w)}(\n_P;E)_w$ for some $w\in W_P$.  Furthermore,
4782: \begin{equation*}
4783: H(\i_P^*\ihat_Q^!\WnC(E))_V \cong \begin{cases}
4784: V[-\l(w)-\#\D_Q] & \text{if $Q = (P,T_V^\eta)$,} \\
4785: 0 & \text{otherwise.}
4786: \end{cases}
4787: \end{equation*}
4788: \end{prop}
4789: 
4790: \begin{proof}
4791: Assume first that $T_V^\eta=P$.  Then we claim that
4792: $H(\i_P^*\ihat_Q^!\WnC(E))_V \cong H(\i_P^*\ihat_Q^!\i_{G*}E)_V$ from which
4793: proposition follows immediately.  To prove the claim, consider the
4794: commutative diagram
4795: \begin{equation*}
4796: \xymatrixnocompile@C-4mm@R-2mm{
4797: {\cdots} \ar[r] &
4798: {H^i(\i_P^*\ihat_Q^!\WnC(E))_V} \ar[r] \ar[d] &
4799: {H^i(\i_P^*\WnC(E))_V} \ar[r] \ar[d] &
4800: {H^i(\i_P^*\jhat_{Q*}\jhat_Q^*\WnC(E))_V} \ar[r] \ar[d] &
4801: {\cdots} \\
4802: {\cdots} \ar[r] &
4803: {H^i(\i_P^*\ihat_Q^!\i_{G*}E)_V} \ar[r] &
4804: {H^i(\i_P^*\i_{G*}E)_V} \ar[r] &
4805: {H^i(\i_P^*\jhat_{Q*}\jhat_Q^*\i_{G*}E)_V} \ar[r] &
4806: {\cdots}
4807: }
4808: \end{equation*}
4809: in which both rows are long exact sequences (obtained from
4810: \eqref{eqnLongCompareLocalCohomologyWithSupports} by setting $Q'=G$ and
4811: then taking $V$-isotypical components).  The middle vertical arrow is an
4812: isomorphism by Lemma
4813: ~\ref{ssectLocalWeightedCohomologyWithAndWithoutProperSupports}.  To show
4814: the right-hand vertical arrow is an isomorphism, consider the corresponding
4815: map on the Mayer-Vietoris spectral sequences from Lemma
4816: ~\ref{ssectRelativeLocalCohomologySupportsSS}.  For
4817: $H(\i_P^*\jhat_{Q*}\jhat_Q^*\WnC(E))_V$ we have
4818: \begin{equation}
4819: \label{eqnWeightedMayerVietoris}
4820: E_1^{p,\cdot} = \bigoplus_{\substack{P < \tilde P \le (P,Q) \\
4821: \#\D_P^{\tilde P} = p+1}} H(\n_P^{\tilde P};H(\i_{\tilde
4822: P}^*\WnC(E)))_V \ .
4823: \end{equation}
4824: The $\tilde P$-term can be rewritten $H(\n_P^{\tilde P};H(\i_{\tilde
4825: P}^*\WnC(E))_{\tilde V})_V$ for an irreducible $L_{\tilde P}$-module
4826: $\tilde V$ satisfying $\xi_{\tilde V}=\xi_V|_{\sa_{\tilde P}}$ (see
4827: \S\ref{ssectKostantDegeneration}).  By Lemma
4828: ~\ref{ssectLocalWeightedCohomologyWithAndWithoutProperSupports} this is
4829: isomorphic to the corresponding term in $E_1$ for
4830: $H(\i_P^*\jhat_{Q*}\jhat_Q^*\i_{G*}E)_V$ provided that $T_{\tilde
4831: V}^\eta=\tilde P$.  But this holds since $T_{\tilde V}^\eta = T_V^\eta \vee
4832: \tilde P$.  The claim then follows from the $5$-lemma.
4833: 
4834: In the case that $T_V^\eta > P$, Lemma
4835: ~\ref{ssectLocalWeightedCohomologyWithAndWithoutProperSupports} shows that
4836: $H(\i_P^*\WnC(E))_V = 0$ and hence $H(\i_P^*\ihat_Q^!\WnC(E))_V \cong
4837: H(\i_P^*\jhat_{Q*}\jhat_Q^*\WnC(E))_V[-1]$.  It also shows that the sum in
4838: \eqref{eqnWeightedMayerVietoris} yields one copy of $V[-\l(w)]$ for each
4839: $\tilde P \in [T_V^\eta, (P,Q)]$.  The differential between two such copies
4840: of $V[-\l(w)]$ for adjacent $\tilde P$ will be an isomorphism by comparison
4841: with the spectral sequence for $H(\i_P^*\jhat_{Q*}\jhat_Q^*\i_{G*}E)_V$.
4842: Consequently $H(\i_P^*\jhat_{Q*}\jhat_Q^*\WnC(E))_V = 0$ unless
4843: $T_V^\eta=(P,Q)$, in which case we obtain $V[-\l(w)-\#\D_Q+1]$.  This
4844: proves the proposition.
4845: \end{proof}
4846: 
4847: \subsection{}
4848: \label{ssectWeightCohomologyMicroSupport}
4849: Let $\eta$ be one of the two {\itshape middle weight profiles\/}:
4850: \begin{equation*}
4851: \begin{aligned}
4852: \text{{\itshape upper middle\/}: } \u &=  - \hsr + \epsilon\hsr \ , \\
4853: \text{{\itshape lower middle\/}: } \v &=  - \hsr \ .
4854: \end{aligned}
4855: \end{equation*}
4856: Here $\epsilon>0$ is chosen sufficiently small such that  $\epsilon\hsr_R <
4857: \chi_R$ for all $R\in\Pl$ maximal, where $\chi_R$ is a positive generator
4858: for $X(S_R^G)$.
4859: 
4860: \begin{thm*}
4861: Let $E$ be an irreducible regular $G$-module and let $\eta$ be
4862: a middle weight profile.  The weak micro-support $\mS_w(\WnC(E))$
4863: consists of those irreducible $L_P$-modules $V$ such that
4864: \begin{enumerate}
4865: \item $V=H^{\l(w)}(\n_P;E)_w$ with $w\in W_P$, and
4866: \label{itemWHMicroPurityNilpotent}
4867: \item $(\xi_V+\hsr)|_{\sa_P^G}=0$.
4868: \label{itemWHMicroPurityVanishing}
4869: \setcounter{saveenum}{\value{enumi}}
4870: \setcounter{saveenumref}{\value{enumi}}
4871: \addtocounter{saveenumref}{-1}
4872: \end{enumerate}
4873: For such $V$ we have
4874: %%%%% Warning: within a \begin{cases} or \matrix, the control sequences \c, \l, and
4875: %%%%% \r are redefined for the \format option.  Thus our definition of \l
4876: %%%%% gets lost.  Use \ell instead.
4877: \begin{equation*}
4878: c(V;\M)=d(V;\M)= \smash[t]{\begin{cases} \ell(w)+\D_P &\qquad \text{if
4879: 			$\eta=\u$,} \\
4880: 			\ell(w)      &\qquad \text{if $\eta=\v$.}
4881: \end{cases}}
4882: \end{equation*}
4883: The micro-support $\mS(\WnC(E))$ consists of those $V\in \mS_w(\WnC(E))$
4884: such that
4885: \begin{enumerate}
4886: \setcounter{enumi}{\value{saveenum}}
4887: \item $V|_{M_P}$ is conjugate self-contragredient\textup;
4888: \label{itemWHMicroPurityTauInvariant}
4889: \end{enumerate}
4890: in this case $w$ from \itemref{itemWHMicroPurityNilpotent} is fundamental.
4891: The micro-support is nonempty if and only if $E|_{\lsp0 G}$ is conjugate
4892: self-contragredient\textup; in this case $\emS(\WnC(E))=\{E\}$.
4893: \end{thm*}
4894: 
4895: \begin{proof}[Proof of Theorem
4896: ~\textup{\ref{ssectWeightCohomologyMicroSupport}}] By Proposition
4897: ~\ref{ssectWeightCohomologyLocalCohomologyWithSupports}, an irreducible
4898: $L_P$-module $V$ is in $\mS(\WnC(E))$ if and only if
4899: \itemref{itemWHMicroPurityNilpotent} and
4900: {\renewcommand{\labelenumi}{(\theenumi)$'$}
4901: \begin{enumerate}
4902: \setcounter{enumi}{\value{saveenumref}}
4903: \item $(P,T_V^\eta)\in[Q_V, Q'_V]$
4904: \label{itemOppositeTInQRange}
4905: \end{enumerate}}
4906: \noindent
4907: hold.  Assume that $\eta=\v=-\hsr$.  Then Condition
4908: \itemref{itemOppositeTInQRange}$'$ is equivalent to
4909: {\renewcommand{\labelenumi}{(\theenumi)$''$}
4910: \begin{enumerate}
4911: \setcounter{enumi}{\value{saveenumref}}
4912: \item $\langle\xi_V+\hsr,\al\spcheck\rangle<0 \quad\Rightarrow\quad
4913: \langle\xi_V+\hsr,\b_\al\spcheck\rangle\ge 0 \quad\Rightarrow\quad 
4914: \langle\xi_V+\hsr,\al\spcheck\rangle \le 0 \quad$
4915: \label{itemOppositeTInQRangeInequalities}
4916: \end{enumerate}}
4917: \noindent
4918: for all $\al\in\D_P$.  For $R\ge P$ define
4919: \begin{equation*}
4920: \begin{split}
4921: \langle \sa_P^*\rangle_R &= \Int{\sa^{*+}_R} - \lsp+\sa_P^{R*} \\
4922: &= \{\, \zeta\in  \sa_P^* \mid \langle\zeta,\g\spcheck\rangle>0 \text{ for
4923: $\g\in\D_R$, and} \\
4924: &\qquad\qquad\qquad\qquad  \langle\zeta,\b_\al^R{}\spcheck\rangle\le 0
4925: \text{ for $\al\in\D_P^R$}\,\}.
4926: \end{split}
4927: \end{equation*}
4928: By Langlands's ``geometric lemmas'' \cite[IV,\S6.11]{refnBorelWallach}
4929: these sets form a disjoint decomposition of $\sa_P^*$, so
4930: \begin{equation}
4931: \xi_V+\hsr_P\in -\langle \sa_P^*\rangle_R \label{eqnWeightInPartition}
4932: \end{equation}
4933: for a unique $R\ge P$.  Now for any $\al\in\D_P\setminus \D_P^R$, write
4934: $\al\spcheck = \al\spcheck_R + \al\spcheck{}^R= \g\spcheck +
4935: \al\spcheck{}^R$, where $\g=\al_R\in \D_R$.  Since $\al\spcheck{}^R \in
4936: -\sa_P^{R+}$, we see from \eqref{eqnWeightInPartition} that
4937: $\langle\xi_V+\hsr,\al\spcheck\rangle<0$.  Thus by the first implication of
4938: \itemref{itemOppositeTInQRangeInequalities}$''$, $(\xi_V+\hsr)_R\in
4939: \lsp+\sa^{*}_R$ which contradicts \eqref{eqnWeightInPartition}.
4940: Consequently we have $R=G$, $\xi_V+\hsr\in \lsp+\sa^{*}_P$, and
4941: $G=(P,T_V^\eta)$.  But then the second implication of
4942: \itemref{itemOppositeTInQRangeInequalities}$''$ implies $\xi_V+\hsr\in -
4943: \sa^{*+}_P$ and thus \itemref{itemWHMicroPurityVanishing} must hold.
4944: Conversely if \itemref{itemWHMicroPurityVanishing} holds, then
4945: \itemref{itemOppositeTInQRangeInequalities}$''$ is automatic.  The fact
4946: that $c(V;\M)=d(V;\M)=\l(w)$ comes from Proposition
4947: ~\ref{ssectWeightCohomologyLocalCohomologyWithSupports}.
4948: 
4949: The case of $\eta=\u$ is similar except that
4950: \itemref{itemOppositeTInQRangeInequalities}$''$ is replaced by
4951: {\renewcommand{\labelenumi}{(\theenumi)$'''$}
4952: \begin{enumerate}
4953: \setcounter{enumi}{\value{saveenumref}}
4954: \item $\langle\xi_V+\hsr,\al\spcheck\rangle<0 \quad\Rightarrow\quad
4955: \langle\xi_V+\hsr,\b_\al\spcheck\rangle> 0 \quad\Rightarrow\quad 
4956: \langle\xi_V+\hsr,\al\spcheck\rangle \le 0 \quad$
4957: \label{itemOppositeTInQRangeInequalitiesUpperWeight}
4958: \end{enumerate}}
4959: \noindent
4960: and we consider the disjoint decomposition of $\sa_P^*$ given by
4961: \begin{equation*}
4962: \begin{split}
4963: \langle\langle \sa_P^*\rangle\rangle_R &= \sa^{*+}_R -
4964: \Int{\lsp+\sa_P^{R*}} \\
4965: &= \{\, \zeta\in  \sa_P^* \mid \langle\zeta,\g\spcheck\rangle\ge0 \text{ for
4966: $\g\in\D_R$, and} \\
4967: &\qquad\qquad\qquad\qquad  \langle\zeta,\b_\al^R{}\spcheck\rangle< 0 \text{
4968: for $\al\in\D_P^R$}\,\}
4969: \end{split}
4970: \end{equation*}
4971: for all $R\ge P$.  We find that
4972: \itemref{itemOppositeTInQRangeInequalitiesUpperWeight}$'''$ is equivalent
4973: to \itemref{itemWHMicroPurityVanishing} and that in this case $\xi_V+\hsr_P
4974: \in -\langle\langle \sa_P^*\rangle\rangle_R$ with $R=P=T_V^\eta$.
4975: 
4976: The assertion that $w$ is fundamental for $V\in\mS(\WnC(E))$ follows from
4977: \itemref{itemWHMicroPurityVanishing} and
4978: \itemref{itemWHMicroPurityTauInvariant} by Lemma
4979: ~\ref{ssectBorelCasselman}.  For the final assertion, note that we can
4980: rephrase our result as $V\in\mS(\WnC(E))$ if and only if $V\preccurlyeq_0
4981: E$ and $V|_{M_P}$ is conjugate self-contragredient.  By Lemma
4982: ~\ref{ssectPartialOrdering} the existence of such a $V$ implies $E|_{\lsp0
4983: G}$ is conjugate self-contragredient.  Also note that if $V\in
4984: \emS(\WnC(E))$ then $Q_V=Q'_V$ by Proposition
4985: ~\ref{ssectWeightCohomologyLocalCohomologyWithSupports} and therefore
4986: $\langle\xi_V+\hsr,\al\spcheck\rangle\neq 0$ for all $\al\in\D_P$.  In view
4987: of \itemref{itemWHMicroPurityVanishing} we must have $P=G$ and thus $V=E$.
4988: \end{proof}
4989: 
4990: \section{Micro-support of Intersection Cohomology}
4991: \label{sectPurity}
4992: In \S\S\ref{sectPurity}, \ref{sectFundamentalCombinatorialPurityProof}, and
4993: \ref{sectPurityProofPartOne}, we consider a {\itshape middle perversity\/}
4994: $p$, either $m(k) = \left\lfloor\frac{k-2}2\right\rfloor$ or $n(k) =
4995: \left\lfloor\frac{k-1}2\right\rfloor$.
4996: 
4997: \begin{thm}
4998: \label{ssectIHMicroPurity}
4999: Assume the irreducible
5000: components of the $\QQ$-root system of $G$ are of type $A_n$, $B_n$,
5001: $C_n$, $BC_n$, or $G_2$.  Let $E$ be an irreducible regular $G$-module
5002: and let $p$ be a middle perversity.  The micro-support $\mS(\IpC(E))$
5003: consists of those irreducible $L_P$-modules $V$ such that
5004: \begin{enumerate}
5005: \item $V=H^{\l(w)}(\n_P;E)_w$ with $w\in W_P$ a fundamental Weyl element
5006: for $P$ in $G$\textup;
5007: \label{itemIHMicroPurityNilpotent}
5008: \item $\langle\xi_V+\hsr,\al\spcheck\rangle \begin{cases} \ge 0 &\qquad
5009: \text{for all $\al\in\D_P$
5010:                       \textup(if $p=m$\textup),}
5011: \\
5012:                       \le 0 &\qquad \text{for all $\al\in\D_P$
5013:                       \textup(if $p=n$\textup)\textup;}
5014: \end{cases}
5015: $
5016: \label{itemIHMicroPurityInequalities}
5017: \item $V|_{M_P}$ is conjugate self-contragredient.
5018: \label{itemIHMicroPurityTauInvariant}
5019: \end{enumerate}
5020: For such a $L_P$-module $V$ we have
5021: %%%%% Warning: within a \begin{cases} or \matrix, the control sequences \c, \l, and
5022: %%%%% \r are redefined for the \format option.  Thus our definition of \l
5023: %%%%% gets lost.  Use \ell instead.
5024: \begin{equation*}
5025: c(V;\M)=d(V;\M)= \begin{cases} \ell(w)+\D_P &\qquad \text{if $p=m$,} \\
5026: 			\ell(w)      &\qquad \text{if $p=n$.}
5027: \end{cases}
5028: \end{equation*}
5029: The essential micro-support $\emS(\IpC(E))$ consists of the maximal elements
5030: of $\mS(\IpC(E))$ under $\preccurlyeq_0$ \textup(see
5031: \textup{\S\ref{ssectPartialOrdering}}\textup)\textup; this is equivalent to
5032: requiring strict inequalities in
5033: \itemref{itemIHMicroPurityInequalities}.
5034: \end{thm}
5035: 
5036: Unlike the case of weighted cohomology, the weak micro-support is not
5037: characterized by simply omitting \itemref{itemIHMicroPurityTauInvariant}.
5038: 
5039: \begin{cor}
5040: \label{ssectIHMicroPurityCorollary}
5041: In the setting of Theorem ~\textup{\ref{ssectIHMicroPurity}} assume that
5042: $E|_{\lsp0 G}$ is conjugate self-contragredient.  Then
5043: $\emS(\IpC(E))=\{E\}$ with $c(E;\IpC(E))= d(E;\IpC(E))= 0$ and
5044: $(\xi_V+\hsr)|_{\sa_P^G}=0$ for all $V\in \mS(\IpC(E))$.
5045: \end{cor}
5046: 
5047: \subsection{}
5048: In the remainder of this section we will reduce the theorem to a
5049: combinatorial vanishing result which will be proven in the following
5050: sections.
5051: 
5052: Recall the ``combinatorial'' intersection cohomology $I_{p_w}H(U)$ (for $U$
5053: an open constructible subset of $|\D_P|$ or $c(|\D_P|)$) defined in
5054: \S\ref{ssectLocalIntersectionCohomology} which depends on $w\in
5055: W$ as well as $p$.  We also consider the version with supports
5056: $I_{p_w}H_Z(U)$ defined for $Z$ constructible and relatively closed in
5057: $U$. This may be interpreted in terms of relative intersection cohomology:
5058: \begin{equation*}
5059: I_{p_w}H_Z(U) = I_{p_w}H(U,U\setminus Z)\ .
5060: \end{equation*}
5061: 
5062: \begin{prop}
5063: \label{ssectLocalIHWithSupports}
5064: Let  $P\le Q\in \Pl$.  Then
5065: \begin{equation*}
5066: H(\i_P^*\ihat_Q^!\IpC(E)) \cong \bigoplus_{w\in
5067: W_P} H(\n_P;E)_w\otimes I_{p_w}H_{c(|\D_P^Q|)}(c(|\D_P|))\ .
5068: \end{equation*}
5069: \end{prop}
5070: \begin{proof}
5071: There is a long exact sequence (set $Q'=G$ in
5072: \eqref{eqnLongCompareLocalCohomologyWithSupports})
5073: \begin{equation*}
5074: \dots \longrightarrow
5075: H^j(\i_P^* \ihat_Q^! \IpC(E)) \longrightarrow H^j(\i_P^* \IpC(E)) \longrightarrow H^j(\i_P^* \jhat_{Q*}\jhat_Q^*
5076: \IpC(E)) \longrightarrow \cdots\ .
5077: \end{equation*}
5078: On the other hand there is a topological long exact sequence
5079: \begin{equation*}
5080: \dots \rightarrow
5081: I_{p_w}H^i_{c(|\D_P^Q|)}(c(|\D_P|)) \rightarrow I_{p_w}H^i(c(|\D_P|)) \rightarrow
5082: I_{p_w}H^i( c(|\D_P|) \setminus c(|\D_P^Q|) ) \rightarrow \cdots\ .
5083: \end{equation*}
5084: We need to show that if the first term of the second sequence is tensored
5085: with $H(\n_{P};E)_w$ and then direct summed over $w\in W_P$, then the
5086: result is isomorphic to the first term of the first sequence.  This is true
5087: for the middle term by Proposition ~\ref{ssectLocalIntersectionCohomology}.
5088: It is also true for the last term: compare the Mayer-Vietoris sequences
5089: abutting to each and apply Proposition
5090: ~\ref{ssectLocalIntersectionCohomology} again.  One may check that these
5091: isomorphisms are compatible with the maps in the long exact sequences.
5092: Application of the $5$-Lemma concludes the proof.
5093: \end{proof}
5094: 
5095: \subsection{}
5096: From the proposition it follows that if an irreducible $L_P$-module $V$
5097: occurs in $\mS(\IpC(E))$, then $V=H^{\l(w)}(\n_P;E)_w$ for some $w\in W_P$,
5098: and this $V$ will actually occur only if $V|_{M_P}$ is conjugate
5099: self-contragredient and $I_{p_w}H_{c(|\D_P^Q|)}(c(|\D_P|) \neq 0$ for some
5100: $Q\in[Q_V,Q'_V]$ (see \S\ref{ssectMicroSupportNotation}).
5101: This condition on $Q$ is equivalent to
5102: \begin{equation}
5103: \begin{aligned}
5104: \langle\xi_V+\hsr,\al\spcheck\rangle \le 0 \qquad &\text{for
5105: $\al\in\Phi(\n_P^Q;\sa_P)$,} \\ 
5106: \langle\xi_V+\hsr,\al\spcheck\rangle \ge 0 \qquad &\text{for
5107: $\al\in\Phi(\n_P^{P,Q};\sa_P)$.}
5108: \end{aligned} \label{eqnWeightInequality}
5109: \end{equation}
5110: 
5111: \subsection{}
5112: \label{ssectSimpleBasicLemma}
5113: Thus we need to show $I_{p_w}H_{c(|\D_P^Q|)}(c(|\D_P|)$ vanishes except for
5114: the cases indicated in Theorem ~\ref{ssectIHMicroPurity}; this is a
5115: combinatorial statement involving the data $\l(w_R)$ for $R\ge P$.  However
5116: our hypotheses give us information on $V$, or equivalently, its highest
5117: weight.  Following an argument from \cite{refnSaperSternTwo}, the simple
5118: basic lemma below will translate the weight information into combinatorial
5119: information; it is a generalization of a result of Casselman
5120: \cite[Prop.~2.6]{refnCasselman} in the real rank one case.  A stronger
5121: result will be presented in \S\ref{sectBasicLemma}.
5122: 
5123: Let $\h=\hb_P+\sa_P$ be a Cartan subalgebra of $\levi_P$ and fix a positive
5124: system $\Phi^+$ of $\Phi(\mathfrak g_\CC,\h_\CC)$ containing
5125: $\Phi(\n_{P\CC},\h_\CC)$.  For $\al\in\Phi(\n_P,\sa_P)$, let
5126: $\n_\al\subseteq \n_P$ be the corresponding root space and write
5127: $\Phi(\n_{\al\CC},\h_\CC)= \{\,\g\in\Phi^+\mid \g|_{\sa_P}= \al\,\}$.  For
5128: $w\in W$, the contribution to $\l(w)$ attributable to $\al$ is $\l_\al(w) =
5129: \#(\Phi(\n_\al,\h_\CC)\cap \Phi_w)$.  Note that $\l_\al(w)=\l_\al(w_P)$ for
5130: $\al\in\Phi(\n_P,\sa_P)$.  Also if $w\in W_P$ then $\l(w) =
5131: \sum_{\al\in\Phi(\n_P,\sa_P)} \l_\al(w)$.
5132: 
5133: \begin{lem*}
5134: Let $P$ be a parabolic $\QQ$-subgroup and let $w\in W_P$.  Let
5135: $V=H^{\l(w)}(\n_P;E)_w$ for an irreducible $G$-module $E$ and assume
5136: $V|_{M_P}$ is conjugate self-contragredient.  For any
5137: $\al\in\Phi(\n_P,\sa_P)$ we have\textup:
5138: \begin{enumerate} 
5139: \item $\langle\xi_V+\hsr,\al\spcheck\rangle \le 
5140: 0 \implies 
5141: \l_\al(w) \ge \frac12\dim \n_\al$.
5142: \label{itemSimpleBasicLemmaLessThan}
5143: \item $\langle\xi_V+\hsr,\al\spcheck\rangle = 0 \implies \l_\al(w) =
5144: \frac12\dim \n_\al$. 
5145: \label{itemSimpleBasicLemmaEqual}
5146: \item $\langle\xi_V+\hsr,\al\spcheck\rangle \ge 0 \implies \l_\al(w) \le 
5147: \frac12\dim \n_\al$.
5148: \label{itemSimpleBasicLemmaGreaterThan}
5149: \end{enumerate}
5150: Furthermore if $\g\in \Phi(\n_\al,\h_\CC)$, then $\Phi_w$ contains
5151: respectively \itemref{itemSimpleBasicLemmaLessThan} at least one,
5152: \itemref{itemSimpleBasicLemmaEqual} exactly one,
5153: \itemref{itemSimpleBasicLemmaGreaterThan} at most one of $\g$ and
5154: $\t'_P\g$.
5155: \end{lem*}
5156: 
5157: \begin{proof}
5158: Let $\lambda\in \h_\CC$ be the highest weight of $E$; by Kostant's theorem
5159: the highest weight of $V$ is then $w(\lambda+\hsr)-\hsr$ and $\xi_V+\hsr =
5160: w(\lambda+\hsr)|_{\sa_P}$.
5161: 
5162: Let $\g\in \Phi(\n_\al,\h_\CC)$.  Since $w(\lambda+\hsr)$ is regular and
5163: $\lambda$ is dominant, $\g\in\Phi_w$ if and only if
5164: $\langle w(\lambda+\hsr),\g\spcheck\rangle \le 0$.  Decompose $\hb_P =
5165: \hb_{P,1}+\hb_{P,-1}$ according to the 
5166: eigenvalues of $\t_P$; by \S\S\ref{ssectConjugateSelfContragredient},~
5167: \ref{ssectConjugateSelfContragredientExample} the hypothesis that
5168: $V|_{M_P}$ is conjugate self-contragredient is equivalent to
5169: $w(\lambda+\hsr)|_{\hb_{P,-1}} = 0$.  On the other hand,
5170: $\g\spcheck|_{\sa_P} = c \al$ where $c>0$.  Thus
5171: \begin{equation*}
5172: \langle w(\lambda+\hsr),\g\spcheck\rangle = \langle
5173: w(\lambda+\hsr)|_{\hb_{P,1}},\g\spcheck\rangle  + c\langle w(\lambda+\hsr),
5174: \al\spcheck\rangle\ .
5175: \end{equation*}
5176: Replacing $\g$ by $\t'_P\g$ in this equation negates the first term and
5177: leaves the second term unchanged, so $\langle
5178: w(\lambda+\hsr),\al\spcheck\rangle \le 0$ implies $\langle
5179: w(\lambda+\hsr),\g\spcheck\rangle \le 0$ or $\langle
5180: w(\lambda+\hsr),\t'_P\g\spcheck\rangle \le 0$.  Thus either $\g$ or
5181: $\t'_P\g$ is in $\Phi_w$.  This proves
5182: \itemref{itemSimpleBasicLemmaLessThan} and the proof of
5183: \itemref{itemSimpleBasicLemmaGreaterThan} is similar;
5184: \itemref{itemSimpleBasicLemmaEqual} follows from
5185: \itemref{itemSimpleBasicLemmaLessThan} and
5186: \itemref{itemSimpleBasicLemmaGreaterThan}.
5187: \end{proof}
5188: 
5189: By \eqref{eqnWeightInequality} and the 
5190: lemma we have
5191: \begin{equation}
5192: \begin{aligned}
5193: \l_\al(w) \ge \tfrac12\dim \n_\al \qquad &\text{for 
5194: $\al\in\Phi(\n_P^Q;\sa_P)$,} \\
5195: \l_\al(w) \le \tfrac12\dim \n_\al \qquad &\text{for 
5196: $\al\in\Phi(\n_P^{P,Q};\sa_P)$.}
5197: \end{aligned} \label{eqnLengthInequality}
5198: \end{equation}
5199: 
5200: \subsection{}
5201: \label{ssectFundamentalCombinatorialPurity}
5202: We begin with a special case of the vanishing theorem.  The proof will 
5203: appear in \S\ref{sectFundamentalCombinatorialPurityProof}; the key point is
5204: that in the case $Q=P$ or $G$ the hypotheses \eqref{eqnLengthInequality} are
5205: preserved under passing by induction from $P$ to $R>P$.
5206: 
5207: \begin{prop*}
5208: Let $p$ be a middle perversity and let $w\in W$.  Let $P\le Q$ be parabolic
5209: subgroups so that \eqref{eqnLengthInequality} is satisfied.  Also assume
5210: that $Q=P$ or $G$.  Then
5211: \begin{equation*}
5212: I_{p_w}H_{c(|\D_{P}^Q|)}(c(|\D_P|))= \begin{cases} 
5213:   \ZZ[-\#\D_P] & \text{if $Q=P$, $p=m$, and
5214:                                  $\l(w_P)=\tfrac12\dim \n_P$,} \\
5215:   \ZZ          & \text{if $Q=G$, $p=n$, and
5216:                                  $\l(w_P)=\tfrac12\dim \n_P$,} \\
5217:   0            & \text{otherwise.}
5218: \end{cases}
5219: \end{equation*}
5220: \end{prop*}
5221: 
5222: \subsection{}
5223: \label{ssectInductionLemma}
5224: For $P<Q<G$ the analogue of \eqref{eqnLengthInequality}
5225: will no longer remain true when we pass by induction to $R>P$.  Nor will
5226: $H(\n_R;E)_{w_R}$ necessarily be conjugate self-contragredient (for any
5227: $E$) so Lemma~\ref{ssectSimpleBasicLemma} will no longer apply directly.  The
5228: following lemma yields an alternate hypothesis that is suitable for
5229: induction and is sufficient.
5230: 
5231: \begin{lem*}
5232: Assume the irreducible components of the $\QQ$-root system of $G$ are of
5233: type $A_n$, $B_n$, $C_n$, $BC_n$, or $G_2$.  Let $P$ be a proper parabolic
5234: subgroup and let $w\in W$.  Let $V=H^{\l(w)}(\n_P;E)_w$ for an irreducible
5235: $G$-module $E$ and assume $V|_{M_P}$ is conjugate self-contragredient.
5236: Assume $Q\ge P$ is a parabolic $\QQ$-subgroup such that
5237: \eqref{eqnWeightInequality} is satisfied.  Then there exists $T\ge P$ so
5238: that
5239: \begin{equation}
5240: 1\le \#\D_T \le 2 \qquad\text{and} \qquad 0\le \#\D_T^{Q\vee T},\!
5241: \ \#(\D_T\setminus \D_T^{Q\vee T}) \le 1 \label{eqnTrestriction}
5242: \end{equation}
5243: and that
5244: \begin{equation}
5245: \begin{aligned}
5246: \l_\al(w) \ge \tfrac12\dim \n_\al \qquad &\text{for
5247: $\al\in\Phi(\n_P;\sa_P)$ such that $\al|_{\sa_T}\in \Phi(\n_T^{Q\vee
5248: T};\sa_T)$,} \\
5249: \l_\al(w) \le \tfrac12\dim \n_\al \qquad &\text{for
5250: $\al\in\Phi(\n_P;\sa_P)$ such that $\al|_{\sa_T}\in\Phi(\n_T^{T,Q\vee
5251: T};\sa_T)$.}
5252: \end{aligned} \label{eqnTLengthInequality}
5253: \end{equation}
5254: Furthermore, if $P<Q<G$ we can assume that
5255: \begin{equation}
5256: \l(w_T) \neq \tfrac12\dim \n_T \text{ when  $\#\D_T=1$.} \label{eqnTNonHalf}
5257: \end{equation}
5258: \end{lem*}
5259: 
5260: The lemma will be proved in \S\ref{sectInductionLemmaProof}.  When $P<Q<G$
5261: this lemma provides the hypotheses for the following theorem:
5262: 
5263: \begin{thm}
5264: \label{ssectCombinatorialPurity}
5265: Let $p$ be a middle perversity and let $w\in W$.  Let $P\le Q$ be parabolic
5266: $\QQ$-subgroups and assume there exists $T\ge P$ satisfying
5267: \eqref{eqnTrestriction}, \eqref{eqnTLengthInequality}, and
5268: \eqref{eqnTNonHalf}.  Then
5269: \begin{equation*}
5270: I_{p_w}H_{c(|\D_{P}^Q|)}(c(|\D_P|))= 0.
5271: \end{equation*}
5272: \end{thm}
5273: 
5274: The proof will appear in \S\ref{sectPurityProofPartOne}.  
5275: 
5276: \subsection{Proof of Theorem ~\ref{ssectIHMicroPurity}}
5277: Note that the nonvanishing cases of Proposition
5278: ~\ref{ssectFundamentalCombinatorialPurity} are relevant only when $Q_V=P$ (for
5279: $p=m$) and $Q_V'=G$ (for $p=n$).  Together with Theorem
5280: ~\ref{ssectCombinatorialPurity}, this implies that an irreducible $L_P$-module
5281: $V$ belongs to $\mS(\IpC(E))$ if and only if
5282: \begin{enumerate}
5283: \renewcommand{\theenumi}{\roman{enumi}}
5284: \renewcommand{\labelenumi}{(\theenumi$'$)}
5285: \item $V=H^{\l(w)}(\n_P;E)_w$ with $w\in W_P$ satisfying $\l(w) =
5286: \frac12 \dim \n_P$;
5287: \label{itemIHMicroPurityNilpotentPrime}
5288: \end{enumerate}
5289: together with \itemref{itemIHMicroPurityTauInvariant} and
5290: \itemref{itemIHMicroPurityInequalities} of Theorem
5291: ~\ref{ssectIHMicroPurity} hold.  The last assertion of
5292: Lemma~\ref{ssectSimpleBasicLemma} implies in this case that $\t_P'$
5293: interchanges $\Phi_w$ and $\Phi(\n_{P\CC},\h_\CC)\setminus \Phi_w$; in
5294: other words, $w$ is a fundamental Weyl element for $P$ in $G$ by Lemma
5295: ~\ref{ssectFundamentalParabolic}.  Thus
5296: \itemref{itemIHMicroPurityNilpotent}$\Longleftrightarrow$(\ref{itemIHMicroPurityNilpotentPrime}$'$)
5297: given \itemref{itemIHMicroPurityTauInvariant} and
5298: \itemref{itemIHMicroPurityInequalities}.  Such a $V$ will lie in
5299: $\emS(\IpC(E))$ furthermore if and only if $Q_V=Q_V'$, that is, if and only
5300: if the inequalities in \itemref{itemIHMicroPurityInequalities} are all
5301: strict.  This is equivalent to $V$ being maximal in $\mS(\IpC(E))$ by Lemma
5302: ~\ref{ssectBorelCasselman}.  \qed
5303: 
5304: \section{Proof of Proposition ~\ref{ssectFundamentalCombinatorialPurity}}
5305: \label{sectFundamentalCombinatorialPurityProof}
5306: 
5307: First assume that $\l(w_P) = \frac12\dim\n_P$.  This implies
5308: $\l(w_R)=\frac12\dim \n_R$ for all $R\ge P$ by
5309: \eqref{eqnLengthInequality} and so
5310: \begin{equation*}
5311: p_w(R)= \begin{cases} \left\lfloor\frac{\#\D_R-2}2\right\rfloor & \text{if
5312:                $p=m$,}\\[2ex]
5313:                \left\lfloor\frac{\#\D_R-1}2\right\rfloor & \text{if $p=n$.}
5314:         \end{cases}
5315: \end{equation*}
5316: For $p=m$ we see that $p_w(R) < \#\D_R-1$ for $R<G$ and thus the link 
5317: cohomology for the $R$-stratum (which inductively is $\ZZ[-\#\D_R 
5318: +1]$) will always be truncated.  For $p=n$ we see that $p_w(R) \ge 0$ 
5319: and thus the link cohomology (which inductively is $\ZZ$) will never 
5320: be truncated.  This proves the proposition in this case.
5321: 
5322: Now assume that $\l(w_P) \neq \frac12\dim\n_P$ and consider the case
5323: $Q=P$.  Let $S> P$ be the unique parabolic $\QQ$-subgroup which is minimal
5324: with the property that $\l(w_S)=\frac12\dim \n_S$.  Consider the Fary
5325: spectral sequence
5326: \begin{equation*}
5327: E_1^{-p,i+p} = \bigoplus_{\#\D_P^R = p}
5328: I_{p_w}H_{|\D_P^R|^\circ}^i(|\D_P|)  \cong
5329: \bigoplus_{\#\D_P^R = p}
5330: I_{p_w}H_{c(\emptyset)}^i(c(|\D_R|)) \Rightarrow 
5331: I_{p_w}H^i(|\D_P|).
5332: \end{equation*}
5333: The hypothesis \eqref{eqnLengthInequality} remains valid if we replace $P$
5334: by $R$ (with $Q=R$) so by induction all the $E_1$ terms vanish except in the
5335: case that $p=m$.  In this case, only the terms for $R\in [S, G]$ are
5336: nonvanishing and
5337: the spectral sequence is isomorphic to the Fary spectral sequence 
5338: abutting to 
5339: $I_{p_w}H^i(c(|\D_S|))$.  This vanishes by the argument in 
5340: the previous paragraph unless $S=G$.  Thus
5341: \begin{equation*}
5342: I_{p_w}H(|\D_P|) = \begin{cases} 0   & \text{if $S < G$ or $p=n$,} \\
5343:                           \ZZ & \text{if $S = G$ and $p=m$.}
5344:                    \end{cases}
5345: \end{equation*}
5346: However in this second case, $p_w(P)
5347: % = \left\lfloor\frac{\dim \n_P - 2\l(w_P) + \#\D_P-2}2\right\rfloor
5348: \ge \left\lfloor\frac{1 +\#\D_P-2}2\right\rfloor \ge 0$ and so a 
5349: degree 0 link cohomology class will not be truncated at the 
5350: $P$-stratum.  Thus $I_{p_w}H_{c(\emptyset)}(c(|\D_P|)) = 0$.
5351: 
5352: The argument for the case $Q=G$ is similar except that we 
5353: use the Mayer-Vietoris spectral sequence converging to $I_{p_w}H^i(|\D_P|)$
5354: from \eqref{eqnp}:
5355: \begin{equation*}
5356: E_1^{p,i-p} =\bigoplus_{\#\D_P^R=p+1} I_{p_w}H^{i-p}(U_R) \cong 
5357: \bigoplus_{\#\D_P^R=p+1} I_{p_w}H^{i-p}_{c(|\D_R|)}(c(|\D_R|))\ . \quad
5358: \qed
5359: \end{equation*}
5360: 
5361: \section{Proof of Lemma ~\ref{ssectInductionLemma}}
5362: \label{sectInductionLemmaProof}
5363: 
5364: We begin with some generalities on chambers in $\sa_P$ and an analogue of
5365: the lemma for the $\QQ$-Weyl group $\lsb \QQ W$ of $G$ (Lemma
5366: ~\ref{ssectQInductionLemma}).  This is a purely combinatorial result
5367: regarding the $\QQ$-root system and its Weyl group.  We then reduce Lemma
5368: ~\ref{ssectInductionLemma} to this result.  Thus $w$ in this section will
5369: refer to an element of $\lsb \QQ W$ until \S\ref{ssectInductionLemmaProof}.
5370: 
5371: \subsection{}
5372: \label{ssectInductionLemmaProofNotation}
5373: Let $\lsb \QQ\Phi\subset \sa^*$ denote the $\QQ$-root system of $G$ with
5374: simple roots $\D$.  Let $P$ be a parabolic $\QQ$-subgroup and let
5375: $\lsb\QQ\Phi^P$ denote the subroot system with basis $\D^P$ in $\sa^{P*}$.
5376: The Weyl group of $\lsb\QQ\Phi^P$ is the parabolic subgroup $\lsb\QQ
5377: W^P\subseteq \lsb\QQ W$ generated by the simple reflections
5378: $\{s_\al\}_{\al\in \D^P}$.  Let $\sb\QQ\Phi_P$ denote the restriction of
5379: the elements$\lsb\QQ\Phi\setminus \lsb\QQ\Phi^P$ to $\sa_P$.  An element of
5380: $\lsb\QQ\Phi_P$ may be expressed as a $\ZZ$-linear combination of the
5381: elements of $\D_P$ with coefficients all $\ge0$ or $\le0$, however in
5382: general $\lsb\QQ\Phi_P$ is not a root system.
5383: 
5384: \subsection{Chambers in $\sa_P$}
5385: \label{ssectSplitComponentChambers}
5386: The ``chambers'' of $\sa_P$
5387: are the connected components of $\sa_P\setminus \bigcup_{\al\in\lsb\QQ\Phi_P}
5388: \ker \al$.  Even though $\lsb\QQ\Phi_P$ is not a root system there is a
5389: combinatorial object that parametrizes these chambers.  Let $\lsb\QQ
5390: W_P$ denote the minimal length representatives of the cosets in
5391: $\lsb\QQ W^P\backslash \lsb\QQ W$ and define%
5392: \begin{equation}
5393: \lsb\QQ W(P) = \{\, w\in \lsb\QQ W\mid w^{-1} \D^P \subseteq \D\,\}
5394: \subseteq \lsb\QQ W_P
5395: \label{eqnSplitWeyl}
5396: \end{equation}
5397: (cf. \cite[I.1.7]{refnMoeglinWaldspurger}%
5398: %
5399: \footnote{Our notation differs in several ways from
5400: \cite{refnMoeglinWaldspurger}.  For one thing, the meaning of
5401: $\lsb\QQ W_P$ and $\lsb\QQ W^P$ are reversed.  Furthermore, since we
5402: consider right cosets of $\lsb\QQ W^P$ rather than left cosets, we
5403: have $w^{-1}$ where \cite{refnMoeglinWaldspurger} would have $w$.}%
5404: %
5405: ).  This is not a subgroup in general.
5406: 
5407: There is a nice geometric way to think of $\lsb\QQ W_P$ and $\lsb\QQ W(P)$.
5408: Namely let $C$ denote the strictly dominant cone of $\lsb\QQ\Phi$ in $\sa$
5409: and let $C^P$ denote the strictly dominant cone of $\lsb\QQ\Phi^P$ in
5410: $\sa^P$.  Then $w\in \lsb\QQ W_P$ if and only if the chamber $wC$ projects
5411: into $C^P$ under the projection $\sa\to \sa^P$.  Furthermore such an
5412: element belongs to $\lsb\QQ W(P)$ if and only if the closure of $wC$ is
5413: adjacent to $\sa_P$ in the sense that $w\overline C \cap \sa_P$ is a
5414: boundary stratum of $w\overline C$ with codimension $\# \D^P$.  In this
5415: case we let $w\cdot C_P \equiv \Int_{\sa_P}(w\overline{C}\cap \sa_P)$
5416: denote the corresponding chamber of $\sa_P$.  (Here $\Int_{\sa_P}$ denotes
5417: the interior of a subset of $\sa_P$.)
5418: 
5419: The correspondence
5420: \begin{equation}
5421: w \quad \longleftrightarrow \quad  w\cdot C_P \label{eqnUnipotentWeylChamber}
5422: \end{equation}
5423: is a bijection of $\lsb\QQ W(P)$ with the chambers of $\sa_P$
5424: \cite[I.1.10]{refnMoeglinWaldspurger}.
5425: 
5426: \subsection{An Alternative Characterization of $\lsb\QQ
5427: W(P)$}
5428: \label{ssectAllOrNone}
5429: Let $\pr_P$ denote the map $\lsb\QQ\Phi\setminus \lsb\QQ\Phi^P \to
5430: \lsb\QQ\Phi_P$ given by restriction to $\sa_P$.  For $w\in \lsb\QQ W$ let
5431: $\lsb\QQ\Phi_w = \{\g \in\lsb\QQ\Phi^+| w^{-1}\g<0\}$.
5432: 
5433: \begin{lem*}
5434: $\lsb\QQ W(P) = \{\, w\in \lsb\QQ W_P \mid \lsb\QQ\Phi_w\cap\pr_P^{-1}(\al)
5435: = \pr_P^{-1}(\al) \text{ or } \emptyset \text{ for all
5436: }\al\in\lsb\QQ\Phi_P^+\,\}.$ \textup(The cases $\pr_P^{-1}(\al)$ and
5437: $\emptyset$ corresponding respectively to $\langle \al, w\cdot
5438: C_P\rangle<0$ and $\langle \al, w\cdot C_P \rangle>0$.\textup)
5439: \end{lem*}
5440: 
5441: \begin{proof}
5442: If $w\in \lsb\QQ W(P)$ and $\g\in \pr_P^{-1}(\al)$ we have
5443: $\g\in\lsb\QQ\Phi_w\Leftrightarrow \langle \g, wC \rangle <0 \Leftrightarrow
5444: \langle \g, w\overline{C}\rangle \le 0 \Leftrightarrow \langle \al, w\cdot
5445: C_P \rangle <0$.  Conversely,
5446: if $w\in \lsb\QQ W_P$ and $\lsb\QQ\Phi_w\cap\pr_P^{-1}(\al) =
5447: \pr_P^{-1}(\al)$ or $\emptyset$ for all $\al\in\lsb\QQ\Phi_P^+$, consider
5448: $\g\in \D^P$.  If $w^{-1}\g=\d_1+\d_2$ ($\d_1$ and $\d_2$
5449: positive roots) is not simple, then $w\d_1>0$ (say) and $w\d_2<0$.
5450: Thus $-w\d_2\in\lsb\QQ\Phi_w$ and $-w\d_2+\g \notin \lsb\QQ\Phi_w$,
5451: while $\pr_P(-w\d_2) = \pr_P(-w\d_2+\g)$.  This contradiction
5452: implies $w^{-1}\D^P\subset \D$ and thus $w\in \lsb\QQ W(P)$.
5453: \end{proof}
5454: 
5455: \begin{lem}
5456: \label{ssectQInductionLemma}
5457: Assume the irreducible components of the $\QQ$-root system of $G$ are of
5458: type $A_n$, $B_n$, $C_n$, $BC_n$, or $G_2$.  Let $P$ be a proper parabolic
5459: $\QQ$-subgroup and let $w\in \lsb\QQ W(P)$.  Then there exists $T\ge P$
5460: such that one of the following hold\textup:
5461: \begin{description}
5462: \item[Case 1] $\#\D_T= 1$.  In this case
5463: $\lsb\QQ\Phi_w\cap\pr_T^{-1}(\lsb\QQ\Phi_T^+) = \pr_P^{-1}(\lsb\QQ\Phi_T^+)$ or $\emptyset$.
5464: \item[Case 2] $\#\D_T= 2$.  In this case $\D_T=\{\b_-,\b_+\}$ and
5465: \begin{equation*}
5466: \begin{aligned}
5467: \lsb\QQ\Phi_w\cap\pr_T^{-1}(\NN \b_-) &= \pr_T^{-1}(\NN \b_-),\\
5468: \lsb\QQ\Phi_w\cap\pr_T^{-1}(\NN \b_+) &= \emptyset.
5469: \end{aligned}
5470: \end{equation*}
5471: \end{description}
5472: \end{lem}
5473: \begin{rem*}
5474: The lemma does not quite imply that
5475: $w_T \in \lsb\QQ W(T)$ since in Case 2 nothing is asserted regarding
5476: $\lsb\QQ\Phi_w\cap\pr_T^{-1}(\b)$ when $\b\in\lsb\QQ\Phi_T^+$ is not a multiple of a
5477: simple root.
5478: \end{rem*}
5479: 
5480: \begin{proof}
5481: It suffices to assume that $\lsb\QQ\Phi$ is irreducible and
5482: even reduced (the case $BC_n$ follows from the case $B_n$).  Let
5483: $\widetilde G$ be the simply connected semisimple $\QQ$-split group with
5484: root system $\lsb\QQ\Phi$.  There always exists an irreducible
5485: representation $V$ of $\widetilde G$ whose nonzero weights form a single
5486: Weyl orbit \cite[\S2]{refnLakshmibaiMusiliSeshadriIII},
5487: \cite{refnKempf}.  We partially order the weights $\Phi(V)$ as usual:
5488: $\u\prec \v$ if $\v-\u$ is a sum of nonnegative integral multiples of
5489: elements of $\D$.  Then the root systems $A_n$, $B_n$, $C_n$, and $G_2$ are
5490: precisely those for which we can arrange that $\Phi(V)$ is totally ordered.
5491: (In the notation of the appendices of \cite{refnBourbakiLie} we choose $V$
5492: to have highest weight $\omega_1$.)  Write $\Phi(V) = \{\e_1\succ
5493: e_2\succ\dots\succ \e_N\}$.  Since the representation is faithful every
5494: root $\g\in \lsb\QQ\Phi^+$ may be expressed (nonuniquely) as $\e_i-\e_j$,
5495: where $1\le i< j\le N$.  Moreover, if $\al\in \D$ is in the support of $\g$
5496: and we represent $\al$ as $\e_k-\e_{k+1}$ for a fixed $k$, we may choose $i$
5497: and $j$ so that $i\le k <j$.  This can be proven in general but is easily
5498: checked for the cases at hand.  Note that to simplify the exposition that
5499: follows we will treat $\e_i-\e_j$ as if it were always a root; any
5500: assertion regarding $\e_i-\e_j$ should be ignored if it is not a root.  For
5501: $w\in \lsb\QQ W$ we will also denote by $w$ the corresponding permutation
5502: of $\{1,2,\dots,N\}$; thus $w(\e_i-\e_j)=\e_{w(i)}-\e_{w(j)}$.
5503: 
5504: First assume that $w(1)> 1$.  Then
5505: \begin{equation}
5506: \begin{aligned}
5507: \e_i - \e_{w(1)}	&\in \lsb\QQ\Phi_w	&& \text{for } i<w(1), \\
5508: \e_{w(1)} - \e_k	&\notin \lsb\QQ\Phi_w	&& \text{for } w(1)<k.
5509: \end{aligned} \label{eqnNegativeColumnPositiveRow}
5510: \end{equation}
5511: Since $w\in \lsb\QQ W_P$ it follows that $\e_{w(1)-1}-\e_{w(1)}\notin
5512: \D^P$.  If $\e_{w(1)}-\e_{w(1)+1}$, \dots, $\e_{N-1}-\e_N\in
5513: \D^P$, then $\pr_P(\e_i-\e_j)=\pr_P(\e_i- \e_{w(1)})$ for all $i<w(1)\le j$
5514: and so by Lemma~\ref{ssectAllOrNone} and \eqref{eqnNegativeColumnPositiveRow} such
5515: $\e_i-\e_j$ must belong to $\lsb\QQ\Phi_w$.  We can thus let $T$ have type
5516: $\D\setminus\{\e_{w(1)-1}-\e_{w(1)}\}$ and Case 1 is satisfied.  Otherwise
5517: let $b\ge w(1)$ be the least index such that $\e_b-\e_{b+1}\notin \D^P$.
5518: Then Lemma~\ref{ssectAllOrNone} and \eqref{eqnNegativeColumnPositiveRow} imply that
5519: \begin{equation*}
5520: \begin{aligned}
5521: \e_i-\e_j	&\in \lsb\QQ\Phi_w	&& \text{for } i< w(1)\le j\le b, \\
5522: \e_j-\e_k	&\notin \lsb\QQ\Phi_w	&& \text{for } w(1)\le j \le b< k.
5523: \end{aligned}
5524: \end{equation*}
5525: Thus we can let $T$ have type
5526: $\D\setminus\{\e_{w(1)-1}-\e_{w(1)},\e_b-\e_{b+1}\}$ and Case 2 is satisfied.
5527: 
5528: Now assume that $w(1)=1$.  If $w(i)=i$ for all $i$ then
5529: $\lsb\QQ\Phi_w=\emptyset$ and letting $T$ have type $\D\setminus
5530: \{\e_i-\e_{i+1}\}$ for any $\e_i-\e_{i+1}\notin \D^P$ will do.  Otherwise
5531: let $a>1$ be the least integer such that $w(a)>a$.  Since
5532: $\e_a-\e_{w(a)}\in \lsb\QQ\Phi_w$ whereas $\e_{a-1}-\e_{w(a)}\notin
5533: \lsb\QQ\Phi_w$ we must have $\e_{a-1}-\e_a\notin \D^P$ by
5534: Lemma~\ref{ssectAllOrNone}.  Furthermore $\e_i-\e_j\notin \lsb\QQ\Phi_w$ for
5535: $i<a\le j$ so we can let $T$ have type $\D\setminus\{\e_{a-1}-\e_a\}$.
5536: \end{proof}
5537: 
5538: \subsection{Proof of Lemma ~\ref{ssectInductionLemma}}
5539: \label{ssectInductionLemmaProof}
5540: We return to the setting of Lemma ~\ref{ssectInductionLemma}; thus $w$ is
5541: now an element of $W$.  Let $\lambda\in\h_\CC$ be the highest weight of
5542: $E$; as in the proof of Lemma ~\ref{ssectSimpleBasicLemma} the hypotheses
5543: on $V$ and $Q$ imply that $\lambda$ satisfies
5544: \begin{equation}
5545: \begin{aligned}
5546: \langle w(\lambda+\hsr),\al\spcheck \rangle  \le 0 \qquad &\text{for 
5547: $\al\in\Phi(\n_P^Q;\sa_P)$,} \\
5548: \langle w(\lambda+\hsr),\al\spcheck \rangle  \ge 0 \qquad &\text{for 
5549: $\al\in\Phi(\n_P^{P,Q};\sa_P)$.}
5550: \end{aligned} \label{eqnWeightInequalityAgain}
5551: \end{equation}
5552: and
5553: \begin{equation*}
5554: \lambda\in
5555: 	B= \{\,\lambda \text{ dominant}\mid
5556: 		w(\lambda+\hsr)|_{\hb_{P,-1}}=0\,\}
5557: \end{equation*}
5558: (see \eqref{eqnEquivalentExpression} and
5559: \eqref{eqnDeltaConjugateSelfContragredient}).  The transformation
5560: $\lambda\mapsto c\lambda+(c-1)\hsr$ for $c\ge 1$ preserves these conditions
5561: (since it corresponds to rescaling $w(\lambda+\hsr)$ by $c$) so we may
5562: assume that $\lambda$ is strictly dominant and hence in the interior of
5563: $B$.  By perturbing $\lambda$ within $B$ slightly we can arrange that all
5564: the inequalities in \eqref{eqnWeightInequalityAgain} are strict
5565: inequalities and that in addition that $\langle w(\lambda+\hsr),\al\spcheck \rangle \neq 0$ for
5566: all $\al\in \Phi(\n_P,\sa_P)$.
5567: 
5568: Since we have arranged that $w(\lambda+\hsr)|_{\sa_P}$ lies in a chamber of
5569: $\sa_P$, the discussion of \S\ref{ssectSplitComponentChambers} and in
5570: particular \eqref{eqnUnipotentWeylChamber} shows that it determines an element
5571: $\lsb\QQ w\in \lsb\QQ W(P)$.  By \eqref{eqnWeightInequalityAgain} and
5572: Lemma~\ref{ssectAllOrNone} we have
5573: \begin{equation}
5574: \begin{aligned}
5575: \Phi_{\lsb\QQ w}\cap\pr_P^{-1}(\al) &= \pr_P^{-1}(\al)
5576: && \Longleftrightarrow &\langle w(\lambda+\hsr),\al\spcheck \rangle  &< 0
5577: && \Longleftarrow  &&\al\in\Phi(\n_P^Q;\sa_P), \\ 
5578: \Phi_{\lsb\QQ w}\cap\pr_P^{-1}(\al) &= \emptyset
5579: && \Longleftrightarrow & \langle w(\lambda+\hsr),\al\spcheck \rangle  &> 0
5580: && \Longleftarrow  &&\al\in\Phi(\n_P^{P,Q};\sa_P). \\ 
5581: \end{aligned} \label{eqnAllOrNoneVersusWeightInequality}
5582: \end{equation}
5583: for all $\al\in\Phi(\n_P,\sa_P)$.  Apply Lemma ~\ref{ssectQInductionLemma}.
5584: If $\#\D_T=1$, denote the unique element of $\D_T$ by $\b_-$ or $\b_+$
5585: depending on whether $\Phi_w\cap\pr_T^{-1}(\Phi_T^+) =
5586: \pr_P^{-1}(\Phi_T^+)$ or $\emptyset$.   In general let $\al_\pm\in \D_P$
5587: restrict to $\b_\pm$.  The lemma and \eqref{eqnAllOrNoneVersusWeightInequality}
5588: imply that $\al_-\in \D_P^Q$ and $\al_+\in \D_P\setminus \D_P^Q$ (which
5589: establishes \eqref{eqnTrestriction}) and furthermore that for any
5590: $\al\in\Phi(\n_P,\sa_P)$,
5591: \begin{equation}
5592: \begin{aligned}
5593: \langle w(\lambda+\hsr),\al\spcheck \rangle  < 0 \qquad &\text{if } \al|_{\sa_T}\in \NN\b_-, \\
5594: \langle w(\lambda+\hsr),\al\spcheck \rangle  > 0 \qquad &\text{if } \al|_{\sa_T}\in \NN\b_+.
5595: \end{aligned} \label{eqnTWeightInequality}
5596: \end{equation}
5597: Lemma ~\ref{ssectSimpleBasicLemma} concludes the proof of \eqref{eqnTLengthInequality}.
5598: 
5599: The final assertion may be proved by induction.  Suppose that $P<Q<G$,
5600: $\D_T=1$ and $\l(w_T)=\tfrac12\dim \n_T$.  Then $\l_\al(w)=\tfrac12 \dim
5601: \n_\al$ for all $\al\in\Phi(\n_T,\sa_P)$.  If $P=Q\cap T$, then
5602: $\D_P^T=\D_P\setminus \D_P^Q$ and so $\l_\al(w) \le \tfrac12 \dim \n_\al$
5603: for all $\al\in\Phi(\n_P^T,\sa_P)$ by Lemma ~\ref{ssectSimpleBasicLemma}.
5604: Then $\widetilde T$ with $\D_P^{\widetilde T}=\D_P^T\setminus \{\al\}$ for
5605: any $\al\in\D_P^T$ has the desired properties.  An analogous argument
5606: applies if $Q\cap T= T$.  Otherwise we can apply the lemma by induction to
5607: the parabolic $\QQ$-subgroups $P/N_T <(Q\cap T)/N_T < T/N_T$ with $w$ and
5608: $\lambda$ replaced by $w^T$ and $w_T(\lambda+\hsr)-\hsr$ respectively to
5609: obtain a parabolic $\QQ$-subgroup $\widetilde T/N_T$.  Then $\widetilde T$
5610: has the desired properties.  \qed
5611: 
5612: \section{Proof of Theorem 
5613: ~\ref{ssectCombinatorialPurity}}
5614: \label{sectPurityProofPartOne}
5615: 
5616: If $T=P$ and $Q=P$ or $G$ we can apply Proposition
5617: ~\ref{ssectFundamentalCombinatorialPurity} to obtain the desired
5618: vanishing; the hypothesis \eqref{eqnLengthInequality} follows from
5619: \eqref{eqnTLengthInequality} and the nonvanishing cases are excluded by
5620: \eqref{eqnTNonHalf}.  So in this section we can assume that if $T=P$, then
5621: $P<Q<G$; this is necessary for Lemma ~\ref{ssectACLemmaOne} below.
5622: 
5623: \renewcommand{\IH}{IH}
5624: \subsection{Notation}
5625: \label{ssectNotationPurityProofOne}
5626: In order to simplify the
5627: notation, we will simply write $\IH(U)$ for $I_{p_w}H(U)$.  Furthermore, if
5628: $Z\subseteq |\D_P|$ is a locally closed constructible subset, we write
5629: \begin{equation*}
5630: \IH_Z = \IH_Z(U)
5631: \end{equation*}
5632: where $U\subseteq |\D_P|$ is any open constructible subset containing $Z$
5633: as a relatively closed subset.  This is well-defined by excision.
5634: 
5635: \subsection{}
5636: Consider the commutative diagram
5637: \begin{equation*}
5638: \xymatrixnocompile@C-3mm@R-2mm{
5639: {\cdots} \ar[r] & {\IH_{c(|\D_P^Q|)}^i(c(|\D_P|))} \ar[r]
5640: & {\IH^i(c(|\D_P|))} \ar[r]^-{r} \ar[d]_{s} &
5641: {\IH^i(c(|\D_P|)\setminus c(|\D_P^Q|))} \ar[r] \ar[d]_{\cong} & {\cdots} \\
5642: {\cdots} \ar[r] &
5643: \composite{i{\IH_{c(|\D_P^Q|)}^i(c(|\D_P|))}*{\IH^i_{|\D_P^Q|}}} \ar[r] &
5644: \composite{i{\IH^i(c(|\D_P|))}*{\IH^i(|\D_P|)}} \ar[r] &
5645: \composite{i{\IH^i(c(|\D_P|)\setminus c(|\D_P^Q|))}*{\IH^i(|\D_P|\setminus |\D_P^Q|)}} \ar[r] & {\cdots}
5646: }
5647: \end{equation*}
5648: in which both rows are long exact sequences.  The vanishing assertion
5649: of Theorem ~\ref{ssectCombinatorialPurity} is equivalent to $r$ being an
5650: isomorphism.  However by the local characterization of intersection
5651: cohomology, the map $s$ is an isomorphism in degrees $i \le p_w(P)$ and the
5652: inclusion of $0$ in degrees $i > p_w(P)$.  Thus we need to show that
5653: \begin{subequations}
5654: \begin{align}
5655: &\IH^{i-1}(|\D_P|\setminus |\D_P^Q|) = 0  &&\text{for $i >
5656: p_w(P) + 1$,} \label{eqnPurityVanishingA} \\
5657: &\Im \left(\IH^{p_w(P)}(|\D_P|\setminus |\D_P^Q|)
5658:       \longrightarrow \IH_{|\D_P^Q|}^{p_w(P)+1}\right) = 0
5659:       &&\text{for $i =
5660: p_w(P) + 1$,} \label{eqnPurityVanishingB} \\
5661: &\IH_{|\D_P^Q|}^i = 0 &&\text{for $i < p_w(P)+1$.}
5662: \label{eqnPurityVanishingC}
5663: \end{align}
5664: \end{subequations}
5665: 
5666: To prove \eqref{eqnPurityVanishingA} we will use the long exact sequence
5667: \begin{equation}
5668: \dots
5669: \longrightarrow \IH^{i-1}_{|\D_P^{Q\vee T}|\setminus |\D_P^Q|}
5670: \longrightarrow \IH^{i-1}(|\D_P|\setminus |\D_P^Q|)
5671: \longrightarrow \IH^{i-1}(|\D_P|\setminus |\D_P^{Q\vee T}|)
5672: \longrightarrow \dots \label{eqng}
5673: \end{equation}
5674: and show the outside terms vanish in the appropriate degrees.
5675: Equation \eqref{eqnPurityVanishingC} will follow similarly using
5676: \begin{equation}
5677: \dots
5678: \longrightarrow \IH^i_{ |\D_P^Q\setminus\D_P^{Q\cap T}|}
5679: \longrightarrow \IH^i_{|\D_P^Q|}
5680: \longrightarrow \IH^i_{|\D_P^Q|\setminus |\D_P^Q\setminus\D_P^{Q\cap T}|}
5681: \longrightarrow \dots \ . \label{eqnDualSequence}
5682: \end{equation}
5683: For \eqref{eqnPurityVanishingB} we will use these arguments to  show that
5684: either $\IH^{p_w(P)}(|\D_P|\setminus |\D_P^Q|)=0$ or
5685: $\IH_{|\D_P^Q|}^{p_w(P)+1} = 0$.
5686: 
5687: \begin{lem}
5688: \label{ssectALemmaTwo}
5689: $\IH^{i-1}_{|\D_P^{Q\vee T}|\setminus |\D_P^Q|}=0$ for all $i$.
5690: \end{lem}
5691: 
5692: \begin{proof}
5693: For $R> P$, let $U_R$ be the star neighborhood of the open face
5694: $|\D_P^R|^\circ$.  Decompose
5695: \begin{equation*}
5696: |\D_P^{Q\vee T}|\setminus |\D_P^Q| =
5697: \coprod_{\emptyset \neq \D_P^R \subseteq \D_P^T\setminus\D_P^{Q \cap T}}
5698:            U_R\cap|\D_P^{Q\vee R}|.
5699: \end{equation*}
5700: The corresponding Fary spectral sequence abutting to
5701: $\IH^{i-1}_{|\D_P^{Q\vee T}|\setminus |\D_P^Q|}$ has (compare
5702: Lemma~\ref{ssectRelativeLocalCohomologySupportsSS})
5703: \begin{equation*}
5704: E_1^{-p,i-1+p} =
5705: \bigoplus_{\substack{\emptyset \neq \D_P^R \subseteq 
5706:                    \D_P^T\setminus\D_P^{Q \cap T}
5707: \\ \#\D_P^{R} = p}}
5708: \negmedspace \IH^{i-1}_{U_R\cap|\D_P^{Q\vee R}|} \cong
5709: \bigoplus_{\substack{\emptyset \neq \D_P^R \subseteq
5710: \D_P^T\setminus\D_P^{Q \cap T} \\
5711: \#\D_P^{R} = p}}
5712: \negmedspace\IH^{i-1}_{c(|\D_{R}^{Q\vee R}|)}(c(|\D_{R}|)) \ .
5713: \end{equation*}
5714: Theorem ~\ref{ssectCombinatorialPurity} may be applied by induction (with $P\le
5715: Q$ replaced by $R\le Q\vee R$ and $T$ remaining the same) to prove that
5716: this vanishes.
5717: \end{proof}
5718: 
5719: Similarly we have the
5720: 
5721: \begin{lem}
5722: \label{ssectCLemmaTwo}
5723: $\IH^i_{|\D_P^Q|\setminus |\D_P^Q\setminus\D_P^{Q\cap T}|} = 0$ for 
5724: all $i$.
5725: \end{lem}
5726: 
5727: \begin{proof}
5728: Cover $|\D_P^Q|\setminus |\D_P^Q\setminus\D_P^{Q\cap T}|$ by 
5729: $\{U_\al\}_{\al\in \D_P^{Q\cap T}}$, where $U_\al$ is the open star of
5730: the vertex $\al$.  The corresponding Mayer-Vietoris 
5731: spectral sequence abutting to $\IH^i_{|\D_P^Q|\setminus 
5732: |\D_P^Q\setminus\D_P^{Q\cap T}|}$ has
5733: (compare Lemma~\ref{ssectRelativeLocalCohomologySupportsSS})
5734: \begin{equation*}
5735: E_1^{p,i-p} =
5736: \bigoplus_{\substack{\emptyset \neq \D_P^R \subseteq \D_P^{Q\cap T}
5737: \\ \#\D_P^{R} = p+1}}
5738: IH^{i-p}_{U_R\cap|\D_P^Q|} \cong
5739: \bigoplus_{\substack{\emptyset \neq \D_P^R \subseteq \D_P^{Q\cap T}
5740: \\ \#\D_P^{R} = p+1}}
5741: \IH^{i-p}_{c(|\D_{R}^{Q}|)}(c(|\D_{R}|)) \ .
5742: \end{equation*}
5743: Theorem ~\ref{ssectCombinatorialPurity} may be applied by induction (with $P\le
5744: Q$ replaced by $R\le Q$ and $T$ remaining the same) to prove that
5745: this vanishes.
5746: \end{proof}
5747: 
5748: \subsection{}
5749: \label{ssectPerversityShift}
5750: In view of the preceding two lemmas, Lemma ~\ref{ssectACLemmaOne} below
5751: will conclude the proof of the theorem.  First we need a useful formula.
5752: For a perversity $p$ let
5753: \begin{equation*}
5754: dp(k) = p(k+1)-p(k)\qquad \text{and} \qquad \ModTwo(k) = 
5755:         \begin{cases} 1 &
5756: 	\text{if $k$ odd,} \\
5757: 	0 & \text{if $k$ even,}
5758: 	\end{cases}
5759: \end{equation*}
5760: Thus for the middle perversities we have $dm=\ModTwo$ and $dn=1-\ModTwo$.
5761: 
5762: \begin{lem*}
5763: For $p$ a middle perversity,
5764: $w\in W$, and $P\le R$,
5765: \begin{equation*}
5766: \begin{split}
5767: p_w(P) &= p_w(R) +
5768: \left( 
5769: 	\tfrac12(\dim\n_P^R+\#\D_P^R) - 
5770: 		\l(w_P^R)
5771: \right)  \\
5772: &\qquad +
5773: \ModTwo(\dim\n_P^R+\#\D_P^R)\left(\tfrac12-dp(\dim\n_P+\#\D_P)\right).
5774: \end{split}
5775: \end{equation*}
5776: \end{lem*}
5777: 
5778: \begin{proof}
5779: This is a simple verification using
5780: $\l(w_P)=\l(w_P^R)+\l(w_R)$ and the definition of a middle perversity.
5781: \end{proof}
5782: 
5783: \begin{lem}
5784: \label{ssectACLemmaOne}
5785: In addition to the hypotheses of Theorem
5786: ~\textup{\ref{ssectCombinatorialPurity}}, assume that if $T=P$, then
5787: $P<Q<G$.  Then the following vanishing results hold\textup:
5788: \begin{enumerate}
5789: \renewcommand{\theenumi}{\alph{enumi}}
5790: \renewcommand{\labelenumi}{(\theenumi)}
5791: \item
5792: $\IH^{i-1}(|\D_P|\setminus |\D_P^{Q\vee T}|)=0 \text{ for }
5793: 	{\begin{cases} i > p_w(P)+1\text{, or} \\
5794: 		i= p_w(P)+1\text{ and} \\
5795: 		\qquad dp(\dim\n_P+\#\D_P)=0.
5796: 	\end{cases}} $
5797: \label{itemACLemmaA}
5798: \smallskip
5799: \addtocounter{enumi}{1}
5800: \item
5801: $\IH^i_{ |\D_P^Q\setminus\D_P^{Q\cap T}|}=0    \text{ for }
5802: 	{\begin{cases} i= p_w(P)+1\text{ and }dp(\dim\n_P+\#\D_P)=1\text{, or}\\
5803: 		i < p_w(P)+1.
5804: 	\end{cases}} $
5805: \label{itemACLemmaC}
5806: \end{enumerate}
5807: \end{lem}
5808: 
5809: \begin{proof}
5810: Consider part \itemref{itemACLemmaA}.  Either $\D_P \setminus \D_P^{Q\vee T}$ is empty
5811: (in which case the lemma is trivial) or by \eqref{eqnTrestriction} it has
5812: just one element.  Let $R=(P,Q\vee T)$ be the corresponding parabolic
5813: subgroup with $\#\D_P^R=1$.  We know that $R\neq G$ since otherwise
5814: $Q\vee T=P$ and hence $Q=T=P$.  Thus $\IH^{i-1}(|\D_P|\setminus
5815: |\D_P^{Q\vee T}|) \cong
5816: \IH^{i-1}(c(|\D_R|))$ will be zero due to truncation at the cone point if
5817: $i-1 > p_w(R)$.  By Lemma ~\ref{ssectPerversityShift} this can be re-expressed
5818: as
5819: \begin{equation}
5820: \begin{split}
5821: &\bigl[ i-(p_w(P)+1) \bigr] +
5822: \bigl[ \tfrac12(\dim\n_P^R+1) - \l(w_P^R) \bigr] \\
5823: &\qquad +
5824: \bigl[ \ModTwo(\dim\n_P^R+1)\left(\tfrac12-dp(\dim\n_P+\#\D_P)\right) \bigr]
5825: > 0.
5826: \end{split} \label{eqnNonvanishingRange}
5827: \end{equation}
5828: Since $\l(w_P^R)\le \tfrac12\dim\n_P^R$ by \eqref{eqnTLengthInequality}, the
5829: second bracketed term of \eqref{eqnNonvanishingRange} is at least $\frac12$.
5830: We are considering $i\ge p_w(P)+1$ so the first term is nonnegative
5831: (and at least $1$ if $i> p_w(P)+1$), while the third term is at least
5832: $-\tfrac12$ (and nonnegative if $dp(\dim\n_P+\#\D_P)=0$).  This proves
5833: part \itemref{itemACLemmaA}.
5834: 
5835: Part \itemref{itemACLemmaC} is similar.  Let $R$ have type $\D_P^R =
5836: \D_P^Q\setminus\D_P^{Q\cap T}= \D_P^{Q\vee T}\setminus \D_P^T$.  Again
5837: $R\neq G$ since otherwise $Q=G$ and $T=P$.  Then $\IH^i_ {
5838: |\D_P^Q\setminus\D_P^{Q\cap T}|}\cong \IH^i_{c(\emptyset)}(c(|\D_{R}|))$
5839: will be zero if $i-1 \le p_w(R)$.  The application of Lemma
5840: ~\ref{ssectPerversityShift} and the inequality $\l(w_P^R)\ge
5841: \tfrac12\dim\n_P^R$ from \eqref{eqnTLengthInequality} concludes the proof.
5842: \end{proof}
5843: 
5844: \specialsection*{Part IV. Satake Compactifications and Functoriality of
5845: Micro-support}
5846: 
5847: This part of the paper begins with two independent sections.  In
5848: \S\ref{sectSatakeCompactifications} we introduce Satake compactifications
5849: $\Xstar_\sigma$ of $X$ and recall Zucker's result
5850: \cite{refnZuckerSatakeCompactifications} that there is a natural quotient
5851: map $\pi\colon  \Xhat\to \Xstar_\sigma$.  In \S\ref{sectEqualRankMicropurityNEW}
5852: we discuss in a fairly general context the functorial behavior of
5853: micro-support of an $\L_W$-module $\M$ under $k^*$ and $k^!$, where
5854: $k\colon Z\hookrightarrow W$ is an inclusion of admissible spaces.  Following
5855: this we proceed to our main result, Theorem~
5856: \ref{ssectRestrictMicroSupportToFiber}, which is a significant
5857: strengthening of this functoriality in the case of restriction to the
5858: fibers of $\pi\colon \Xhat\to \Xstar_\sigma$.
5859: 
5860: \section{Satake Compactifications}
5861: \label{sectSatakeCompactifications}
5862: We will briefly outline the theory of Satake compactifications
5863: $\Xstar_\sigma$ and give Zucker's realization
5864: \cite{refnZuckerSatakeCompactifications} of $\Xstar_\sigma$ as a quotient
5865: space of $\Xhat$.  References are \cite{refnSatakeCompact},
5866: \cite{refnSatakeQuotientCompact}, \cite{refnBorelEnsemblesFondamentaux},
5867: \cite{refnZuckerSatakeCompactifications}, and
5868: \cite{refnCasselmanGeometricRationality}.
5869: 
5870: \subsection{Notation}
5871: Let $S\subseteq \lsb\RR S$ be maximal tori of $G$ split over $\QQ$ and
5872: $\RR$ respectively and choose compatible orderings on the roots.  For
5873: $k=\QQ$ or $\RR$ let $\lsb k\D$ denote the simple restricted $k$-roots; as
5874: usual we omit the left subscript when $k=\QQ$.  We will be working with a
5875: regular representation $\sigma\colon G \to GL(U)$.  Let $\u$ denote highest
5876: weight of $\sigma$ and let $\lsb k\u=\u|_{\lsb k\sa}$ be the restricted
5877: highest $k$-weight.  A subset $\Th\subseteq \lsb k\D$ is said to be
5878: {\itshape $\sigma$-connected\/} if $\Th\cup \{\lsb k\u\}$ is connected.
5879: Here a finite subset $\Th$ of an inner product space is said to be
5880: {\itshape connected\/} if the associated graph (with vertices $\Th$ and an
5881: edge between $\al$ and $\b$ whenever $(\al,\b)\neq 0$) is connected.  For
5882: any subset $\Th\subseteq\lsb k\D$, let $\kap(\Th)$ be the largest
5883: $\sigma$-connected subset of $\Th$.  If $\Th$ is a $\sigma$-connected
5884: subset of $\lsb k\D$, let $\o(\Th)$ be the largest subset of $\lsb k\D$
5885: such that $\kap(\o(\Th))=\Th$.  As in
5886: \cite{refnCasselmanGeometricRationality} we extend this notation to
5887: arbitrary subsets $\Th\subseteq \lsb k\D$ by setting
5888: $\o(\Th)=\o(\kap(\Th))$; it is easy to see that $\o(\Th)\supseteq \Th$.
5889: 
5890: If $P$ is a parabolic $\RR$-subgroup, let $P^\dag$ be the parabolic
5891: $\RR$-subgroup containing $P$ with $\RR$-type $\o(\DRR^P)$.  This is the
5892: unique largest parabolic $\RR$-subgroup containing $P$ with
5893: $\kap(\DRR^{P^\dag})=\kap(\DRR^P)$.  In general a parabolic $\RR$-subgroup
5894: will be called {\itshape $\sigma$-saturated\/} if it has $\RR$-type
5895: $\o(\Th)$ for some $\Th\subseteq \DRR$.  Let $\restr\colon \DRR\to \D\cup\{0\}$
5896: denote%
5897: \footnote{Our use in this section of this standard notation should not be
5898: confused with our equally standard use elsewhere of $\hsr$ to denote one-half
5899: the sum of the positive roots.}
5900: %
5901: the restriction map on simple roots.  For a subset $\Th\subseteq \DRR$ set
5902: $\Th\sphat= \restr(\Th)\setminus\{0\}$ and for a subset $\U\subseteq \D$
5903: set $\Utilde= \restr^{-1}(\U\cup\{0\})$.  Subsets of $\DRR$ of the form
5904: $\Utilde$ are called {\itshape $\QQ$-rational\/} and are precisely the
5905: $\RR$-types of the parabolic $\QQ$-subgroups.  Clearly $(\Utilde)\sphat=\U$
5906: and $\widetilde{(\Th\sphat)}$ is the smallest $\QQ$-rational set containing
5907: $\Th$.
5908: 
5909: \subsection{Satake compactifications of $D$}
5910: \label{ssectSatakeCompactificationsSymmetricSpaces}
5911: In this subsection and the next we depart from our usual context and assume
5912: that $G$ is merely defined over $\RR$ with $K\subseteq G(\RR)$ a maximal
5913: compact subgroup and that $D$ denotes the associated symmetric space
5914: $G(\RR)/K\lsb\RR A_G$.
5915: 
5916: Let $\sigma\colon G \to \GL(U)$ be an irreducible regular representation of $G$
5917: which is nontrivial on every simple factor of $G(\RR)$ and fix an
5918: admissible metric $h_0$ on $U$ with respect to $K$ as in
5919: \S\ref{ssectMetrics}.  Let $T^*$ denote the adjoint with respect to $h_0$
5920: of an endomorphism $T$ of $U$ and let $S(U)$ denote the real vector space
5921: of self-adjoint endomorphisms.  The representation $\sigma$ induces a
5922: natural representation $\tilde \sigma\colon G(\RR) \to \GL(\PP S(U))$ given by
5923: \begin{equation*}
5924: \tilde\sigma(g)[T]= [\sigma(g)\circ T \circ\sigma(g)^*]\ .
5925: \end{equation*}
5926: The class of the identity
5927: endomorphism is fixed by $K\lsb\RR A_G$ under $\tilde\sigma$ and thus the
5928: map $g \mapsto [\sigma(g)\sigma(g)^*]$ descends a $G(\RR)$-equivariant map
5929: \begin{equation*}
5930: %\label{eqnSatakeEmbedding}
5931: D \longrightarrow \PP S(U)\ .
5932: \end{equation*}
5933: This map is an embedding; the closure of the image is denoted $\lsb\RR
5934: \Dstar_\sigma$, the {\itshape Satake compactification%
5935: %
5936: \footnote{Here we have followed most closely Zucker
5937: \cite{refnZuckerSatakeCompactifications}; Borel
5938: \cite{refnBorelEnsemblesFondamentaux} uses an action on the right for $U$
5939: and $S(U)$ and hence replaces $\u$ by the lowest weight of $U$.  Satake's
5940: construction \cite{refnSatakeCompact} is identical except that he
5941: represents $S(U)$ by Hermitian matrices.  Casselman
5942: \cite{refnCasselmanGeometricRationality} dispenses with $U$ altogether and
5943: works directly with an irreducible representation $V$ containing a
5944: $K$-fixed vector.  In \cite[\S3]{refnSaperGeometricRationality} it is
5945: shown that given $U$ there exists $V$ (in fact contained in $S(U)_\CC$)
5946: such that Casselman's construction agrees with Satake's construction.  The
5947: highest weight of this $V$ is $2\lsb\RR\u$ and hence the notion of
5948: $\sigma$-connected defined here agrees with that in
5949: \cite{refnCasselmanGeometricRationality} for subsets of $\D$ or $\DRR$.}
5950: %
5951: of $D$ associated to $\sigma$\/}.  The Satake compactification is
5952: independent of the choice of $K$ and $h_0$; usually $\sigma$ will be fixed
5953: and we omit it from the notation.  If $D$ is Hermitian symmetric then the
5954: closure of the Harish-Chandra embedding of $D$ as a bounded domain is
5955: called the {\itshape natural compactification\/} and it is topologically
5956: equivalent to the Satake compactification for a certain $\sigma$.
5957: 
5958: \subsection{Real boundary components}
5959: \label{ssectRealBoundaryComponents}
5960: A {\itshape real Satake boundary component\/} $D_{P,h}\subseteq \lsb\RR
5961: \Dstar$ is the set of points fixed under the action of $N_P(\RR)$ for some
5962: parabolic $\RR$-subgroup $P$.  The action of $P(\RR)$ descends to an action
5963: of $L_P(\RR)$ on $D_{P,h}$.  The centralizer of $D_{P,h}$ under this action
5964: is the group of real points of a normal reductive $\RR$-subgroup
5965: $L'_{P,\l}\subseteq L_P$.  The action then descends further to the real
5966: points of $L'_{P,h} = L_P/L'_{P,\l}$, and thus $D_{P,h}$ is realized as
5967: $L'_{P,h}(\RR)/(K_P\cap L'_{P,h}(\RR))$, the symmetric space associated to
5968: $L'_{P,h}(\RR)$.  (We use the subscripts $h$ and $\l$ to mimic the notation
5969: frequently used in the Hermitian case.)
5970: 
5971: We can describe these groups concretely in terms of roots.  Set
5972: $\Th=\kap(\DRR^P)$; then the orthogonal decomposition $\DRR^P = \Th \coprod
5973: (\DRR^P\setminus\Th)$ yields a decomposition of the Levi factor as an
5974: almost direct product,
5975: \begin{equation}
5976: L_P = \widetilde{L'_{P,h}} L'_{P,\l}. \label{eqnLevisFactors}
5977: \end{equation}
5978: Here $\widetilde{L'_{P,h}}$ is the minimal normal connected semisimple
5979: $\RR$-subgroup with $\Th$ as simple $\RR$-roots; it is a lift of
5980: $L'_{P,h}$.  The group $L'_{P,\l}$ is the almost direct product of the
5981: minimal normal connected semisimple $\RR$-subgroup with
5982: $\DRR^P\setminus\Th$ as simple $\RR$-roots together with the center of
5983: $L_P$ and all almost simple factors of $L_P$ with compact groups of real
5984: points \cite[5.11]{refnBorelTits}.
5985: 
5986: The parabolic $\RR$-subgroup $P$ giving rise to a boundary component
5987: $D_{P,h}$ as above is not unique.  To index $D_{P,h}$ in a unique fashion
5988: we use its normalizer $\{\,g\in G(\RR) \mid gD_{P,h}=D_{P,h}\,\}$; this is
5989: the group of real points of a $\sigma$-saturated parabolic $\RR$-subgroup,
5990: namely $P^\dag$.  Recall that $P^\dag$ is the unique parabolic
5991: $\RR$-subgroup containing $P$ with $\RR$-type $\o(\DRR^P)$.  For later use
5992: we note that the centralizer $\{\,g\in G(\RR) \mid gy=y \text{ for all
5993: $y\in D_{P,h}$}\,\}$ of $D_{P,h}$ is the group of real points of the
5994: inverse image of $L'_{P^\dag,\l}$ under $P^\dag\to L_{P^\dag}$.  On the
5995: other hand, if $R$ is a $\sigma$-saturated parabolic $\RR$-subgroup,
5996: $R(\RR)$ is the normalizer of a unique boundary component $D_{R,h}$, namely
5997: the points fixed by $N_R(\RR)$.  We have
5998: \begin{equation*}
5999: \lsb\RR \Dstar_\sigma = \coprod_{\substack{\text{$R$ $\sigma$-saturated}\\
6000: 				\text{parabolic $\RR$-subgroup}}} D_{R,h}\ .
6001: \end{equation*}
6002: 
6003: \subsection{Rational boundary components}
6004: \label{ssectRationalBoundaryComponents}
6005: Now consider again a connected reductive algebraic $\QQ$-group $G$ with
6006: $D=G(\RR)/KA_G$.  We assume that $\lsb\RR S_G = S_G$.  (This is not
6007: necessary but simplifies the exposition; the assumption is satisfied if $D$
6008: is Hermitian or $\lsp0 G$ is equal-rank.)  Then the preceding two
6009: subsections may be applied to yield a Satake compactification $\lsb\RR
6010: \Dstar_\sigma$.
6011: 
6012: A real boundary component $D_{R,h}$ of $\lsb\RR \Dstar_\sigma$ is called
6013: {\itshape rational\/} \cite[\S\S3.5,~3.6]{refnBailyBorel} if
6014: \begin{enumerate}
6015: \item its normalizer $R$ is defined over $\QQ$, and
6016: \label{itemRationalNormalizer}
6017: \item
6018: \label{itemRationalCentralizer}
6019: the group $L'_{R,\l}$ contains a normal subgroup $L_{R,\l}$ of $L_R$
6020: defined over $\QQ$ such that $L'_{R,\l}(\RR)/L_{R,\l}(\RR)$ is compact.%
6021: %
6022: \footnote{In \cite{refnBorelEnsemblesFondamentaux} it is assumed that
6023: $L'_{R,\l}$ itself is defined over $\QQ$, but it is noted in
6024: \cite{refnBailyBorel} that this is too restrictive.}
6025: %
6026: \end{enumerate}
6027: Here Condition ~\itemref{itemRationalNormalizer} is equivalent to
6028: $\G_{N_R}\back N_R(\RR)$ being compact while Condition
6029: ~\itemref{itemRationalCentralizer} ensures that $D_{R,h}$ may be realized
6030: as the symmetric space associated to the reductive algebraic $\QQ$-group
6031: $L_{R,h}\equiv L_R/L_{R,\l}$.  For a rational boundary component $D_{R,h}$,
6032: equation \eqref{eqnLevisFactors} may be replaced by a factorization by
6033: $\QQ$-subgroups,
6034: \begin{equation}
6035: L_R = \widetilde{L_{R,h}} L_{R,\l}. \label{eqnLevisRationalFactors}
6036: \end{equation}
6037: 
6038: On the other hand, consider the real boundary components $D_{R,h}$ for
6039: which
6040: \begin{equation}
6041: \label{eqnSiegelBoundaryComponents}
6042: \DRR^R = \o(\Utilde) \qquad \text{for some $\sigma$-connected
6043: $\U\subseteq \D$.}
6044: \end{equation}
6045: Real boundary components satisfying \eqref{eqnSiegelBoundaryComponents} are
6046: exactly those in the closure of a Siegel set \cite[Lemma
6047: ~8.1]{refnCasselmanGeometricRationality}.  By \cite[Proposition
6048: ~3.3(i)]{refnZuckerSatakeCompactifications} any real boundary component
6049: satisfying Condition ~\itemref{itemRationalNormalizer} satisfies
6050: \eqref{eqnSiegelBoundaryComponents}.  The Satake compactification $\lsb\RR
6051: \Dstar_\sigma$ is called {\itshape geometrically rational\/} if every real
6052: boundary component satisfying \eqref{eqnSiegelBoundaryComponents} is
6053: rational.%
6054: %
6055: \footnote{Our presentation here follows Casselman
6056: \cite{refnCasselmanGeometricRationality}.  The assumption that Condition
6057: ~\itemref{itemRationalNormalizer} holds for the real boundary components
6058: satisfying \eqref{eqnSiegelBoundaryComponents} is the assertion that
6059: $\o(\Utilde)$ is $\QQ$-rational for any $\sigma$-connected subset
6060: $\U\subseteq \D$.  A weaker assumption is made in
6061: \cite[(3.3)]{refnZuckerSatakeCompactifications} (Assumption 1) and the
6062: above assertion is deduced via \cite[Proposition
6063: ~3.3(ii)]{refnZuckerSatakeCompactifications}.  However as pointed out in
6064: \cite[\S9]{refnCasselmanGeometricRationality}, this proposition is
6065: incorrect.}
6066: %
6067: Examples of geometrically rational Satake compactifications include those
6068: where $\sigma$ is defined over $\QQ$ \cite[Theorem
6069: ~8]{refnSaperGeometricRationality} or where $D$ is a Hermitian symmetric
6070: space and $\lsb\RR \Dstar_\sigma$ is the natural compactification
6071: \cite{refnBailyBorel}.  Casselman \cite{refnCasselmanGeometricRationality}
6072: gives a necessary and sufficient criterion for geometric rationality in
6073: terms of $\lsb\RR\u$ and the Tits index of $G$.
6074: 
6075: For geometrically rational Satake compactifications, Condition
6076: ~\itemref{itemRationalNormalizer} in the definition of rational boundary
6077: component implies Condition ~\itemref{itemRationalCentralizer}.  Thus the
6078: rational boundary components of a geometrically rational Satake
6079: compactification are indexed by the $\sigma$-saturated parabolic
6080: $\QQ$-subgroups.  In fact by the following lemma these are precisely the
6081: parabolic $\QQ$-subgroups with $\QQ$-type $\o(\U)$ for $\U\subseteq \D$.
6082: 
6083: \begin{lem*}
6084: If $\lsb\RR \Dstar_\sigma$ is geometrically rational,
6085: \begin{alignat}{2}
6086: \label{eqnQQsaturatedImpliesRRsaturated}
6087: \widetilde{\o(\U)}&= \o(\Utilde) &\qquad &\text{for $\U\subseteq \D$,} \\
6088: \label{eqnRationalRRsaturatedImpliesQQsaturated}
6089: \o(\Th)\sphat&= \o(\Th\sphat) &\qquad &\text{for $\Th\subseteq \DRR$ with
6090: $\o(\Th)$ $\QQ$-rational.}
6091: \end{alignat}
6092: \end{lem*}
6093: \begin{proof}
6094: Note that a subset $\U\subseteq \D$ is
6095: $\sigma$-connected if and only if there is a $\sigma$-connected subset
6096: $\Th\subseteq \DRR$ with $\Th\sphat = \U$
6097: \cite[(2.4)]{refnZuckerSatakeCompactifications} and that this implies that
6098: \begin{equation}
6099: \label{eqnKapSphatCommute}
6100: \kap(\Th\sphat) = \kap(\Th)\sphat\qquad\text{for any $\Th\subseteq \DRR$}\ .
6101: \end{equation}
6102: To prove \eqref{eqnQQsaturatedImpliesRRsaturated} we first claim that
6103: $\kap(\widetilde{\o(\U)})= \kap(\Utilde)$.  Since $\widetilde{\o(\U)}$
6104: contains $\kap(\Utilde)$ it follows that $\kap(\widetilde{\o(\U)})\supseteq
6105: \kap(\Utilde)$.  On the other hand, $\kap(\widetilde{\o(\U)})\sphat$ is
6106: $\sigma$-connected and contained in $\o(\U)$ and therefore
6107: $\kap(\widetilde{\o(\U)})\sphat \subseteq \kap(\U) \subseteq \U$; it
6108: follows that $\kap(\widetilde{\o(\U)}) \subseteq \kap(\Utilde)$ which
6109: finishes the proof of the claim.  The claim implies that
6110: $\widetilde{\o(\U)}\subseteq \o(\Utilde)$.  For the opposite inclusion it
6111: suffices (since by geometric rationality both sets are $\QQ$-rational) to
6112: check that $\o(\U)\supseteq \o(\Utilde)\sphat$.  This follows since
6113: $\kap(\U) = \kap(\Utilde)\sphat = \kap(\o(\Utilde)\sphat)$ by
6114: \eqref{eqnKapSphatCommute}.  For
6115: \eqref{eqnRationalRRsaturatedImpliesQQsaturated} suppose $\o(\Th)$ is
6116: $\QQ$-rational for $\Th\subseteq \DRR$ and set $\U=\kap(\Th)\sphat$.  By
6117: \cite[Proposition ~3.3(i)]{refnZuckerSatakeCompactifications},
6118: $\kap(\Th)=\kap(\Utilde)$ and hence $\o(\Th)=\o(\Utilde)=
6119: \widetilde{\o(\U)}$ (use \eqref{eqnQQsaturatedImpliesRRsaturated} for the
6120: last equality); we thus see that $\o(\Th)\sphat= \o(\U)$ which in view of
6121: \eqref{eqnKapSphatCommute} completes the proof.
6122: \end{proof}
6123: 
6124: \subsection{Satake compactifications of $X$}
6125: Assume $\lsb\RR \Dstar_\sigma$ is a geometrically rational Satake
6126: compactification of $D$.  Let $\Dstar_\sigma$ be the union of $D$ and the
6127: rational boundary components.  By the preceding discussion we have
6128: \begin{equation*}
6129: \Dstar_\sigma = \coprod_{\substack{\text{$R$ $\sigma$-saturated}\\
6130: 				\text{parabolic $\QQ$-subgroup}}} D_{R,h}
6131: \end{equation*}
6132: and $G(\QQ)$ acts on $\Dstar_\sigma$.  There is a topology on
6133: $\Dstar_\sigma$ (the ``Satake topology'') for which $\Xstar_\sigma \equiv
6134: \G\back \Dstar_\sigma$ is a compact Hausdorff space containing $X$ as a
6135: dense open subset; $\Xstar_\sigma$ is the {\itshape Satake compactification
6136: of $X$ associated to $\sigma$\/} and as usual we will omit $\sigma$ from
6137: the notation.  For example, if  $D$ is a Hermitian symmetric space and
6138: $\lsb\RR \Dstar_\sigma$ is the natural compactification, then $\Xstar$
6139: is the Baily-Borel-Satake compactification \cite{refnBailyBorel}.
6140: 
6141: Let $\Pl^*_\sigma$ denote the $\G$-conjugacy classes of $\sigma$-saturated
6142: parabolic $\QQ$-subgroups.  For $R\in \Pl^*_\sigma$, set $F_R =
6143: \G_{L_{R,h}}\back D_{R,h}$, where we view $D_{R,h}$ as the symmetric space
6144: associated to $L_{R,h}$ as above and $\G_{L_{R,h}} = \G_{L_R}/(\G_{L_R}\cap
6145: L_{R,\l})$.  Then $\Xstar_\sigma$ has a stratification
6146: \begin{equation*}
6147: \Xstar_\sigma = \coprod_{R\in \Pl^*_\sigma} F_R\ .
6148: \end{equation*}
6149: We define a partial order on the rational boundary components (and
6150: similarly on the strata of $\Xstar_\sigma$) by setting $D_{R,h}\le
6151: D_{R',h}$ if and only if $D_{R,h}\subseteq \cl{D_{R',h}}$; this corresponds
6152: to the conditions that $R^\g \cap R'$ is a parabolic $\QQ$-subgroup (for
6153: some $\g\in \G$ and $\kap(\D^R)\subseteq \kap(\D^{R'})$.
6154: 
6155: \subsection{$\Xstar_\sigma$ as a quotient of $\Xhat$}
6156: Continue to assume that $\lsb\RR \Dstar_\sigma$ is a geometrically
6157: rational Satake compactification of $D$.  Let $P$ be a parabolic
6158: $\QQ$-subgroup and let $\DRR^P= \Utilde$ be its $\RR$-type.  By
6159: \eqref{eqnQQsaturatedImpliesRRsaturated}, $\o(\Utilde)$ is
6160: $\QQ$-rational so $P^\dag$ is a $\sigma$-saturated parabolic
6161: $\QQ$-subgroup and thus defines a stratum $D_{P^\dag,h}$ of $\Dstar$.
6162: Let $D_{P,\l} = L_{Q,\l}(\RR)/K_{Q,\l}A_Q$ be the symmetric space of
6163: the second factor of \eqref{eqnLevisRationalFactors} (where $K_{Q,\l} =
6164: K_Q\cap L_{Q,\l}(\RR)$).  The factorization \eqref{eqnLevisRationalFactors}
6165: induces a decomposition
6166: \begin{equation}
6167: D_{P} = L_P(\RR)/(K_PA_P) = D_{P,h}\times D_{P,\l}
6168: \label{eqnEhatsFactors}
6169: \end{equation}
6170: and we let
6171: \begin{equation*}
6172: q_P\colon   D_{P} \to D_{P^\dag,h}
6173: \end{equation*}
6174: be the projection onto the first factor.  (Note that
6175: $D_{P,h}=D_{P^\dag,h}$ since $\kap(\Utilde)=\kap(\o(\Utilde))$.)
6176: Together these maps yield a surjection
6177: \begin{equation*}
6178: q\colon  \widehat D \to \Dstar_\sigma \ .
6179: \end{equation*}
6180: Let $\pi\colon \Xhat\to \Xstar_\sigma$ be the map induced by $q$.  Zucker's
6181: main result in \cite{refnZuckerSatakeCompactifications} is that $\pi$ is a
6182: quotient map.  (Note however that the quotient topology on $\Dstar_\sigma$
6183: induced by $q$ may be finer than the Satake topology on $\Dstar_\sigma$
6184: \cite[(3.10)]{refnZuckerSatakeCompactifications}.)  Note that in the
6185: notation of \S\ref{ssectPullbackToFiber} we have $\pi^{-1}(F_R)=
6186: X_R(L_{R,\l})$ for any $\sigma$-saturated parabolic $\QQ$-subgroup $R$.
6187: 
6188: \section{Functoriality under Inverse Images}
6189: \label{sectEqualRankMicropurityNEW}
6190: Let $k\colon Z\hookrightarrow W$ be an inclusion map of admissible spaces
6191: possessing unique maximal strata and let $\M$ be an $\L_W$-module.  Let
6192: $X_S$ denote the unique maximal stratum of $W$.  We wish to study the
6193: micro-support of the inverse images $k^*\M$ and $k^!\M$.  In order to
6194: preserve the condition of conjugate self-contragredience, we shall
6195: sometimes assume certain boundary components are equal-rank: a symmetric
6196: space is called {\itshape equal-rank\/} if it can be expressed as
6197: $G(\RR)/K$, where $G$ is a reductive $\RR$-group, $K$ is a maximal compact
6198: subgroup of $G(\RR)$, and $G$ is equal-rank (see
6199: \S\ref{ssectEqualRankGroups}).  Thus if $G$ is defined over $\QQ$ and $D$
6200: is defined as usual to be $G(\RR)/KA_G$, then $D$ is equal-rank if and only
6201: if $\lsp0G$ is equal-rank, which is the case if and only if $V|_{\lsp0G}$
6202: is conjugate self-contragredient for any regular $G$-module $V$.
6203: 
6204: \subsection{Notation}
6205: \label{ssectEqualRankMicropurityNotation}
6206: Given an admissible space $W$ with unique maximal stratum $X_S$ and an
6207: irreducible $L_P$-module $V\in \IrrRep(W)$, recall that $Q^W_V$
6208: (resp. $Q^{\prime W}_V$) is the parabolic $\QQ$-subgroup containing $P$
6209: with type $\{\,\al\in \D_P^S \mid
6210: \langle\xi_V+\hsr,\al\spcheck\rangle <0\,\}$ (resp. $\{\,\al\in \D_P^S \mid
6211: \langle \xi_V+\hsr,\al\spcheck\rangle \le 0\,\}$) with respect to $P$.
6212: 
6213: \subsection{The case of an open embedding}
6214: \label{ssectMicroSupportOpenInverseImage}
6215: The following proposition is clear:
6216: 
6217: \begin{prop*} Let $k\colon Z\hookrightarrow W$ be the inclusion of a open
6218: admissible subspace and let $\M$ be an $\L_W$-module.  Then
6219: $\mS(k^*\M)=\mS(\M)\cap\IrrRep(\L_{Z})$.  \textup(Note that $k^*=k^!$
6220: in this case.\textup)
6221: \end{prop*}
6222: 
6223: \subsection{More partial orderings on $\IrrRep(\L_W)$}
6224: \label{ssectAdditionalPartialOrdering}
6225: Before treating the case of a closed embedding, we need to generalize the
6226: partial ordering of \S\ref{ssectPartialOrdering}.  Let $V$, $\tilde V\in
6227: \IrrRep(\L_W)$ be irreducible $L_P$- and $L_{\tilde P}$-modules
6228: respectively.  Let $\lsp+\sa_P^{\tilde P*}$ denote the convex cone
6229: generated by all $\al\in\D_P^{\tilde P}$.  Set $V \preccurlyeq_+ \tilde V$
6230: if the following three conditions are fulfilled:
6231: \begin{enumerate}
6232: \item \label{itemParabolic} $P\le \tilde P$;
6233: \item \label{itemKostantComponent} $V= H^{\l(w)}(\n_{P}^{\tilde P};
6234: \tilde V)_w$ for some  $w\in W_P^{\tilde P}$;
6235: \item \label{itemRootCone} $(\xi_V+\hsr)|_{\sa_P^{\tilde P}} \in
6236: \lsp+\sa_P^{\tilde P*}$.
6237: \end{enumerate}
6238: Similarly set $V \preccurlyeq_- \tilde V$ if instead of
6239: \itemref{itemRootCone} the condition
6240: \begin{equation*}
6241: (\xi_V+\hsr)|_{\sa_P^{\tilde P}} \in -\lsp+\sa_P^{\tilde P*}
6242: \end{equation*}
6243: is fulfilled.  If both $V \preccurlyeq_+ \tilde V$ and $V \preccurlyeq_-
6244: \tilde V$ we recover $V\preccurlyeq_0 \tilde V$ .  If $V\preccurlyeq_\pm
6245: \tilde V$ we set $[\tilde V:V]=\l(w)$ where $w$ is as in
6246: \itemref{itemKostantComponent}.
6247: 
6248: \begin{lem}
6249: \label{ssectPartialOrderingLemma}
6250: Both $\preccurlyeq_+$ and $\preccurlyeq_-$ are partial orderings on
6251: $\IrrRep(\L_W)$.  They are generated by the relations where $\#\D_P^{\tilde
6252: P}=1$.
6253: \end{lem}
6254: 
6255: \begin{proof}
6256: The only issue in the first assertion is transitivity.  This is clear for
6257: \itemref{itemParabolic} and for \itemref{itemKostantComponent} it follows
6258: from \eqref{eqnLinkComposition}.  As for \itemref{itemRootCone}, suppose
6259: $V\preccurlyeq_+ \tilde V \preccurlyeq_+ \tilde{\tilde V}$.  Then
6260: \begin{equation*}
6261: (\xi_V+\hsr)|_{\sa_P^{\tilde{\tilde P}}} = 
6262: (\xi_V+\hsr)|_{\sa_P^{\tilde P}} +
6263: (\xi_{\tilde V}+\hsr)|_{\sa_{\tilde P}^{\tilde{\tilde P}}}
6264: \end{equation*}
6265: so we need to check that the elements of both $\D_P^{\tilde P}$ and
6266: $\D_{\tilde P}^{\tilde{\tilde P}}$ are in $\lsp+\sa_P^{\tilde{\tilde
6267: P}*}$ when viewed as linear functions on $\sa_P^{\tilde{\tilde P}} =
6268: \sa_P^{\tilde P} + \sa_{\tilde P}^{\tilde{\tilde P}}$.  But this is
6269: one of Langlands's ``geometric lemmas'' \cite[IV,
6270: \S6.5(2)]{refnBorelWallach}.
6271: 
6272: For the second assertion, suppose that $V\preccurlyeq_+ \tilde{\tilde V}$.
6273: If $\langle \xi_V+\hsr,\al\spcheck\rangle< 0$ for all $\al \in
6274: \D_P^{\tilde{\tilde P}}$, then $(\xi_V+\hsr)|_{\sa_P^{\tilde{\tilde P}}}$
6275: belongs to the interior of $-\lsp+\sa_P^{\tilde{\tilde P}*}$ \cite[IV,
6276: \S6.2]{refnBorelWallach}, which is a contradiction.  So there exists
6277: $\al_0\in \D_P^{\tilde{\tilde P}}$ such that
6278: $\langle\xi_V+\hsr,\al_0\spcheck\rangle\ge 0$.  Let $\tilde P\ge P$ have
6279: type $\D_P^{\tilde P}=\{\al_0\}$ and decompose $w=w^{\tilde P}w_{\tilde
6280: P}\in W_P^{\tilde P}W_{\tilde P}^{\tilde{\tilde P}}$.  Then it is easy to
6281: check that $\tilde V= H^{\l(w_{\tilde P})}(\n_{\tilde P}^{\tilde{\tilde
6282: P}}; \tilde{\tilde V})_{w_{\tilde P}}$ satisfies $V\preccurlyeq_+ \tilde
6283: V\preccurlyeq_+ \tilde{\tilde V}$ and we can use induction.
6284: \end{proof}
6285: 
6286: \begin{lem}
6287: \label{ssectBehaviorOfQsubVUnderPartialOrder}
6288: If $V\preccurlyeq_+ V'$ then $Q_V^W\vee P' \ge Q_{V'}^W$ and $Q_V^{\prime
6289: W}\vee P' \ge Q_{V'}^{\prime W}$.  If $V\preccurlyeq_- V'$ then $Q_V^W\vee
6290: P' \le Q_{V'}^W$ and $Q_V^{\prime W}\vee P' \le Q_{V'}^{\prime W}$.
6291: \end{lem}
6292: \begin{proof}
6293: By Lemma~\ref{ssectPartialOrderingLemma} it suffices to consider the case
6294: where $\D_P^{P'}=\{\al_0\}$.  Suppose $V\preccurlyeq_+ V'$ and thus
6295: $\langle \xi_V+\hsr,\al_0\spcheck\rangle\ge0$.  Let $\al'\in \D_{P'}$ and
6296: let $\al\in \D_P\setminus \D_P^{P'}$ be the unique element for which
6297: $\al|_{\sa_{P'}}=\al'$.  Since $\xi_{V'} + \hsr_{P'}=
6298: (\xi_V+\hsr)|_{\sa_{P'}}$ we have
6299: \begin{equation*}
6300: \langle \xi_{V'} + \hsr,\al'{}\spcheck\rangle =
6301: \langle \xi_V+\hsr,\al\spcheck\rangle - 
6302: \langle \b_{\al_0}^{P'}, \al\spcheck\rangle
6303: \langle \xi_V+\hsr,\al_0\spcheck\rangle \ .
6304: \end{equation*}
6305: Since $\langle \b_{\al_0}^{P'}, \al\spcheck\rangle = c\langle \al_0,
6306: \al\spcheck\rangle\le 0$ (where $c>0$), the inequality $\langle\xi_{V'} +
6307: \hsr,\al'{}\spcheck\rangle<0$ implies
6308: $\langle\xi_V+\hsr,\al\spcheck\rangle<0$ and thus $Q_V^W\vee P' \ge
6309: Q_{V'}^W$.  The other assertions are proven similarly.
6310: \end{proof}
6311: 
6312: \subsection{The case of a closed embedding: weak form}
6313: \label{ssectWeakMicroSupportClosedInverseImage}
6314: Let $\M$ be an $\L_W$-module.
6315: 
6316: If $k\colon Z\hookrightarrow W$ is the inclusion of an admissible subspace
6317: possessing a unique maximal stratum and  $\mathscr S\subseteq
6318: \IrrRep(\L_W)$ is any subset, define
6319: \begin{equation}
6320: \begin{aligned}
6321: k^*\mathscr S &= \{\,V\in \IrrRep(\L_{Z}) \mid
6322: V\preccurlyeq_+ \tilde V \text{ for some }\tilde V\in\mathscr S\,\} \\
6323: k^!\mathscr S &= \{\,V\in \IrrRep(\L_{Z}) \mid
6324: V\preccurlyeq_- \tilde V \text{ for some }\tilde V\in\mathscr S\,\}\ .
6325: \end{aligned}
6326: \label{eqnRestrictMicroSupportClosedStratum}
6327: \end{equation}
6328: 
6329: \begin{prop*}
6330: Let $\ihat_R\colon \Xhat_R\cap W\hookrightarrow W$ be the inclusion of a closed
6331: stratum and let $\M$ be an $\L_W$-module.  Then
6332: \begin{equation*}
6333: \mS_w(\ihat_R^*\M) \subseteq \ihat_R^*\mS_w(\M) \qquad
6334: \text{and}\qquad
6335: \mS_w(\ihat_R^!\M) \subseteq \ihat_R^!\mS_w(\M)\ .
6336: \end{equation*}
6337: For $V\in \mS_w(\ihat_R^*\M)$ we have estimates
6338: \begin{equation}
6339: \begin{aligned}
6340: c(V;\ihat_R^*\M) &\ge \min_{V \preccurlyeq_+ \tilde V} ( c(\tilde V;\M) +
6341: [\tilde V:V] ) \\ 
6342: d(V;\ihat_R^*\M) &\le
6343: \max_{V \preccurlyeq_+ \tilde V} ( d(\tilde V;\M) + [\tilde V:V] )\ .
6344: \end{aligned}
6345: \label{eqnWeakMSInverseImageDegree}
6346: \end{equation}
6347: For $V\in \mS_w(\ihat_R^!\M)$ we have estimates
6348: \begin{equation}
6349: \begin{aligned}
6350: c(V;\ihat_R^!\M) &\ge \min_{V \preccurlyeq_- \tilde V} ( c(\tilde V;\M) +
6351: [\tilde V:V] + \dim \sa_P^{\tilde P}) \\ 
6352: d(V;\ihat_R^!\M) &\le \max_{V \preccurlyeq_- \tilde V} ( d(\tilde V;\M) +
6353: [\tilde V:V] + \dim \sa_P^{\tilde P})\ .
6354: \end{aligned}
6355: \label{eqnWeakMSInverseImageProperSupportDegree}
6356: \end{equation}
6357: \end{prop*}
6358: 
6359: \begin{proof}
6360: We need to show for all $R\in\Pl(W)$ that
6361: \begin{equation}
6362: \label{eqnInductionClaim}
6363: \parbox{.8\linewidth}{if $V\in \mS_w(\ihat_R^*\M)$ then there exists
6364: $\tilde V\in\mS_w(\M)$ with $V\preccurlyeq_+ \tilde V$}
6365: \end{equation}
6366: and
6367: \begin{equation}
6368: \label{eqnInductionClaimDegree}
6369: c(\tilde V;\M) + [\tilde V:V] \le c(V;\ihat_R^*\M)\le d(V;\ihat_R^*\M) \le
6370: d(\tilde V;\M) + [\tilde V:V]\ ,
6371: \end{equation}
6372: and the analogous claims for $\mS_w(\ihat_R^!\M)$.
6373: 
6374: First some notation.  Let $P\in \Pl(\Xhat_R\cap W)$ and let $V$ be an
6375: irreducible $L_P$-module.  Let $Q\in[Q_V^R, Q_V^{\prime
6376: R}]$.  (Here we write $Q_V^R$ instead of $Q_V^{\Xhat_R\cap W}$ and
6377: similarly for $Q_V^{\prime R}$.)  Set $U=(P,R)\cap S$ and let
6378: $T=(Q,R)\cap S= Q\vee U$:
6379: \begin{equation*}
6380: \xymatrix @ur @M=1pt @R=1.5pc @C=1pc {
6381: {U} \ar@{-}[r] \ar@{-}[d] & {Q_V^R\vee U} \ar@{-}[r] \ar@{-}[d] & {T}
6382: \ar@{-}[r] \ar@{-}[d] & {Q_V^{\prime R}\vee U} \ar@{-}[r] \ar@{-}[d] & {S} \ar@{-}[d] \\
6383: {P} \ar@{-}[r] & {Q_V^R} \ar@{-}[r] & {Q} \ar@{-}[r]
6384: & {Q_V^{\prime R}} \ar@{-}[r] & {R}
6385: }
6386: \end{equation*}
6387: By Proposition
6388: ~\ref{ssectFunctorsOnLsheaves}\itemref{itemFunctorsComposition}\itemref{itemFunctorsSquare}
6389: we have $\i_P^* \ihat_Q^!\ihat_R^*\M = \i_P^* \ihat_{T}^!\M$ and $\i_P^*
6390: \ihat_Q^!\ihat_R^!\M = \i_P^* \ihat_{Q}^!\M$.  Thus the condition that
6391: $V\in \mS_w(\ihat_R^*\M)$ becomes
6392: \begin{equation}
6393: \label{eqnStratumStarMicroSupport}
6394: H^j(\i_P^* \ihat_T^!\M)_V\neq 0
6395: \end{equation}
6396: for some $j$ and some $Q$ as above, and the condition that $V\in
6397: \mS_w(\ihat_R^!\M)$ becomes
6398: \begin{equation}
6399: \label{eqnStratumShriekMicroSupport}
6400: H^j(\i_P^* \ihat_Q^!\M)_V\neq 0\ .
6401: \end{equation}
6402: 
6403: The following case of \eqref{eqnInductionClaim} is easy and will be
6404: used below:
6405: \begin{equation}
6406: \label{eqnAllowableSignStarCase}
6407: \parbox{.8\linewidth}{if $V\in \mS_w(\ihat_R^*\M)$ and $Q_V^{\prime W} =
6408: Q_V^{\prime R}\vee U$, then $V\in\mS_w(\M)$.}
6409: \end{equation}
6410: For $Q_V^W\in [Q_V^R, Q_V^R\vee U]$ and thus $T\in [Q_V^W, Q_V^{\prime
6411: W}]$.  Hence \eqref{eqnStratumStarMicroSupport} implies that $V\in
6412: \mS_w(\M)$.  Similarly 
6413: \begin{equation}
6414: \label{eqnAllowableSignShriekCase}
6415: \parbox{.8\linewidth}{if $V\in \mS_w(\ihat_R^!\M)$ and $Q_V^{W} =
6416: Q_V^{R}$, then $V\in\mS_w(\M)$.}
6417: \end{equation}
6418: 
6419: For simplicity we now consider the general case in detail only for
6420: $V\in \mS_w(\ihat_R^*\M)$; the changes necessary for $V\in
6421: \mS_w(\ihat_R^!\M)$ will be indicated at the end.  We will use
6422: induction on $\#\D_P^S$.  Since the case where $\#\D_P^S=0$ is
6423: trivial, we may assume that \eqref{eqnInductionClaim} and
6424: \eqref{eqnInductionClaimDegree} are true for all $R\in\Pl(W)$ and all
6425: irreducible $L_{P'}$-modules $V'$ satisfying $\#\D_{P'}^S < \#\D_P^S$.
6426: 
6427: Let $P'\ge P$ have type $\D_P^{P'}= \{\,\al\in \D_P^S\setminus \D_P^R
6428: \mid \langle\xi_V+\hsr,\al\spcheck\rangle > 0 \,\}$ and let $T'\le T$ and $U'\le U$ have
6429: types $\D_P^{T'} = \D_P^T\setminus \D_P^{P'}$ and $\D_P^{U'} =
6430: \D_P^U\setminus \D_P^{P'}$ respectively:
6431: \begin{equation*}
6432: \xymatrix @ur @M=1pt @R=1.5pc @C=1pc {
6433: {} & {U} \ar@{-}[rr] \ar@{-}[dd] & {}  & {T}
6434: \ar@{-}[rr] \ar@{-}[dl] \ar@{-}[dd] & {}  & {S} \ar@{-}[dd] \\
6435: {U'} \ar@{-}'[r][rr] \ar@{-}[ru] \ar@{-}[dd] & {}  & {T'} \ar@{-}'[r][rr]
6436: \ar@{-}'[d][dd] & {} & {R\vee U'} \ar@{-}[ru] \ar@{-}'[d][dd] \\
6437: {} & {P'} \ar@{-}[rr] & {}  & {Q\vee P'}
6438: \ar@{-}[rr] \ar@{-}[dl] & {}  & {R\vee P'} \\
6439: {P} \ar@{-}[rr] \ar@{-}[ru] & {}  & {Q} \ar@{-}[rr]
6440: & {} & {R} \ar@{-}[ru]
6441: }
6442: \end{equation*}
6443: We can assume that $P'>P$ since otherwise $V\in\mS_w(\M)$ by
6444: \eqref{eqnAllowableSignStarCase}.  Consider the short exact sequence
6445: \begin{equation*}
6446: 0\to \i_P^* \ihat_{T'}^!  \M \to \i_P^* \ihat_{T}^! \M \to
6447: \i_P^* \i_{P'*}(\i_{P'}^*\ihat_{T}^!  \M) \to 0
6448: \end{equation*}
6449: and corresponding long exact sequence (which is the analogue of the long
6450: exact sequence of a triple)
6451: \begin{equation}
6452: \dots\to
6453: H^j(\i_P^* \ihat_{T'}^!  \M)_V \to
6454: H^j(\i_P^* \ihat_{T}^! \M)_V \to
6455: H^j(\i_P^* \i_{P'*}(\i_{P'}^*\ihat_{T}^!  \M))_V \to \cdots\ .
6456: \label{eqnExactTripleTransfer}
6457: \end{equation}
6458: Since by \eqref{eqnStratumStarMicroSupport} the middle term is
6459: nonzero, either the first or last term must be nonzero as well.  If
6460: the first term of \eqref{eqnExactTripleTransfer} is nonzero, then we
6461: have similarly to \eqref{eqnAllowableSignStarCase} that
6462: $V\in\mS_w(\M)$ (since $Q_V^W \le Q_V^R\vee U' \le Q\vee U' = T' \le
6463: Q_V^{\prime R} \vee U' =Q_V^{\prime W}$ in the current situation).  If
6464: instead the last term of \eqref{eqnExactTripleTransfer} is nonzero, it
6465: follows from \eqref{eqnPushForwardType} that
6466: $H^\l(\n_P^{P'};H^{j-\l}(\i_{P'}^*\ihat_{T}^!\M))_V \neq 0$.  Thus
6467: \begin{equation}
6468: H^{j-\l}(\i_{P'}^*\ihat_{T}^!\M)_{V'}\neq 0
6469: \label{eqnPPrimeStarTShriekNonvanishing}
6470: \end{equation}
6471: for some irreducible $L_{P'}$-module $V'$ satisfying $V\preccurlyeq_+ V'$
6472: and $\l=[V':V]$.
6473: 
6474: Choose $R'\in [P',S]\cap \Pl(W)$ such that $Q'\equiv R'\cap T \le
6475: Q_{V'}^{\prime R'} $.  (For example, one can set $Q'= Q_{V'}^{\prime W}\cap
6476: T$ and take $R'=(Q',T)\cap S'$.  Another choice of $R'$ will be made in
6477: \S\ref{sectFunctorialityMicroSupportEqualRank}.)  Let $S' = R' \vee T$\/:
6478: \begin{equation*}
6479: \xymatrix @ur @M=1pt @R=1.5pc @C=1pc {
6480: {} & {T}
6481: \ar@{-}[r] \ar@{-}[d] &  {S'} \ar@{-}[r] \ar@{-}[d] & {S} \\
6482: {P'} \ar@{-}[r] & {Q'} \ar@{-}[r]
6483: & {R'}
6484: }
6485: \end{equation*}
6486: By Proposition
6487: ~\ref{ssectFunctorsOnLsheaves}\itemref{itemFunctorsComposition}\itemref{itemFunctorsSquare},
6488: $\i_{P'}^*\ihat_{T}^!\M = \i_{P'}^* \ihat_{Q'}^!\ihat_{R'}^*\ihat_{S'}^!\M$
6489: and so \eqref{eqnPPrimeStarTShriekNonvanishing} becomes
6490: \begin{equation}
6491: H^{j-\l}(\i_{P'}^* \ihat_{Q'}^!(\ihat_{R'}^*\ihat_{S'}^!\M))_{V'}\neq 0\ .
6492: \label{eqnVPrimeInTerm}
6493: \end{equation}
6494: Furthermore we claim that
6495: \begin{equation}
6496: Q_{V'}^{R'}\le Q' \le Q_{V'}^{\prime R'}\ .
6497: \label{eqnQPrimeIsAllowable}
6498: \end{equation}
6499: The second inequality of \eqref{eqnQPrimeIsAllowable} is our
6500: hypothesis on $Q'$; the first inequality of \eqref{eqnQPrimeIsAllowable}
6501: follows since
6502: \begin{equation*}
6503: Q' = R'\cap T = R'\cap (Q\vee U) \ge R'\cap (Q_V^R\vee U) \ge Q_V^{R'}\vee
6504: P' \ge Q_{V'}^{R'}\ ,
6505: \end{equation*}
6506: where at the last stage we use Lemma
6507: ~\ref{ssectBehaviorOfQsubVUnderPartialOrder}.
6508: 
6509: Equations \eqref{eqnVPrimeInTerm} and \eqref{eqnQPrimeIsAllowable} show
6510: that $V'\in\mS_w(\ihat_{R'}^*\ihat_{S'}^!\M)$.  By the inductive
6511: hypothesis we can apply \eqref{eqnInductionClaim} to $V'$ and conclude
6512: there exists an irreducible $L_{\tilde P}$-module $\tilde V\in
6513: \mS_w(\ihat_{S'}^!\M)$ with $V' \preccurlyeq_+ \tilde V$.  By transitivity
6514: of $\preccurlyeq_+$ (Lemma ~\ref{ssectPartialOrderingLemma}) we have
6515: $V\preccurlyeq_+ \tilde V$.  However $S' \ge T\vee \tilde P \ge Q_V^W \vee
6516: \tilde P \ge Q_{\tilde V}^W$ by Lemma
6517: ~\ref{ssectBehaviorOfQsubVUnderPartialOrder}, and thus $\tilde V\in
6518: \mS_w(\M)$ by \eqref{eqnAllowableSignShriekCase}.  This proves
6519: \eqref{eqnInductionClaim}.  Equation \eqref{eqnInductionClaimDegree} also
6520: follows since $[\tilde V: V] = [\tilde V: V'] + [V': V]$.
6521: 
6522: The proof for $\mS_w(\ihat_R^!\M)$ is similar except that we let
6523: $P'\ge P$ have type $\D_P^{P'}= \{\,\al\in \D_P^S\setminus \D_P^R \mid
6524: \langle\xi_V+\hsr,\al\spcheck\rangle < 0 \,\}$ and use the long exact sequence
6525: \begin{equation}
6526: \dots\to
6527: H^j(\i_P^* \ihat_{Q}^!  \M)_V \to
6528: H^j(\i_P^* \ihat_{Q\vee P'}^! \M)_V \to
6529: H^j(\i_P^* \i_{P'*}(\i_{P'}^*\ihat_{Q\vee P'}^!  \M))_V \to \cdots\ .
6530: \end{equation}
6531: It is now the first term that is nonzero and hence either the middle
6532: or last term (in degree $j-1$) must be nonzero as well.  Furthermore
6533: we choose $R'$ for which $Q'\equiv R'\cap (Q\vee P')\ge Q_{V'}^{R'}$ and let
6534: $S'=(Q',Q\vee P')\cap S$\/:
6535: \begin{equation*}
6536: \xymatrix @ur @M=1pt @R=1.5pc @C=1pc {
6537: {} & {Q\vee P'}
6538: \ar@{-}[rr] \ar@{-}[d] &  {} & {S} \ar@{-}[d] \\
6539: {P'} \ar@{-}[r] & {Q'} \ar@{-}[r]
6540: & {R'} \ar@{-}[r] & {S'}
6541: }
6542: \end{equation*}
6543: (For example we can take $R'=(Q\vee P',Q_{V'}^W \vee Q\vee P')\cap S$ for which
6544: $Q'=Q\vee P'$ and $S'=S$.)
6545: Equation  \eqref{eqnVPrimeInTerm} must be replaced by
6546: \begin{equation*}
6547: H^{j-1-\l}(\i_{P'}^*\ihat_{Q'}^!
6548: 	(\ihat_{R'}^!\ihat_{S'}^*\M))_{V'}\neq 0\ .\qed
6549: \end{equation*}
6550: \renewcommand{\qed}{}
6551: \end{proof}
6552: 
6553: \begin{rem}
6554: The proposition is false in general if $\mS_w$ is replaced by $\mS$.  We
6555: will see in \S\ref{sectFunctorialityMicroSupportEqualRank} a situation in
6556: which this replacement is valid.
6557: \end{rem}
6558: 
6559: \subsection{The case of an inclusion of a fiber}
6560: \label{ssectMicroSupportInclusionFiber}
6561: Say we are given a connected normal $\QQ$-subgroup $L_{R,\l}\subseteq L_R$
6562: for a given $R\in\Pl$ and set $L_{R,h}=L_R/L_{R,\l}$; we have a almost
6563: direct product decomposition $L_R= \widetilde{L_{R,h}}L_{R,\l}$ where $
6564: \widetilde{L_{R,h}}$ is a lift of $L_{R,h}$.  Let $X_R(L_{R,\l})\subseteq
6565: \Xhat_R$ be the partial compactification of $X_R$ in the
6566: ``$L_{R,\l}$-directions'' as in \S\ref{ssectPullbackToFiber}.  Let
6567: $\ihat_{R,\l}\colon \Xhat_{R,\l}\hookrightarrow X_R(L_{R,\l})$ be the inclusion
6568: of a generic fiber of the flat bundle $\pi\colon  X_R(L_{R,\l})\to X_{R,h}$.
6569: 
6570: Note that any $V\in \IrrRep(L_R)$ may be written as $\tilde V_h\otimes
6571: V_\l$, where $\tilde V_h\in\IrrRep(\widetilde{L_{R,h}})$ and
6572: $V_\l\in\IrrRep(L_{R,\l})$; the restriction $\Res_{L_{P,\l}}^{L_P} V$ is a
6573: sum of copies of $V_\l$.  For a subset $\mathscr S \subseteq
6574: \IrrRep(\L_{X_R(L_{R,\l})})$, define
6575: \begin{multline}
6576: \Res_\l^*\mathscr S = \Res_\l^!\mathscr S = \\
6577: \biggl\{\,V_\l\in
6578: \IrrRep(\L_{\Xhat_{R,\l}}) \biggm|
6579: \parbox[c]{.5\linewidth}{$V_\l$ is a component of
6580: $\Res_{L_{P,\l}}^{L_P} V$ for some $L_P$-module $V\in \mathscr
6581: S$}\,\biggr\}\ .
6582: \label{eqnRestrictMicroSupportRellFiber}
6583: \end{multline}
6584: 
6585: \begin{prop*}
6586: Let $\M$ be an $\L_{X_R(L_{R,\l})}$-module and assume that $X_{R,h}$ is
6587: equal-rank.  Then $\mS(\ihat_{R,\l}^*\M) = \Res_\l^* \mS(\M)$ and for
6588: $V_\l\in \mS(\ihat_{R,\l}^*\M)$ we have the estimate
6589: \begin{equation}
6590: [c(V_\l;\ihat_{R,\l}^*\M),d(V_\l;\ihat_{R,\l}^*\M)] = \bigcup_{
6591: \Res_\l^*V=V_\l} [c(V;\M),d(V;\M)]\ .
6592: \label{eqnDegreeFiberRestrictionEstimates}
6593: \end{equation}
6594: The same result holds for $\ihat_{R,\l}^!\M$ except for the shift in degree
6595: of $-\dim D_{R,h}$.
6596: \end{prop*}
6597: 
6598: \begin{proof}
6599: For $P_\l\le Q_\l\in \Pl(\Xhat_{R,\l})$ we compute via
6600: \eqref{eqnPullbackToFiberFormulas} that
6601: \begin{equation}
6602: \i_{P_\l}^*\ihat_{Q_\l}^!(\ihat_{R,\l}^*\M) =
6603: \i_{P_\l}^*\ihat_{P,\l}^*(\i_P^*\ihat_Q^!\M)\ ,
6604: \label{eqnMicroSupportOnFiberCalculation}
6605: \end{equation}
6606: where $P\le Q\in \Pl(X_R(L_{R,\l}))$ are defined as in
6607: \S\ref{ssectPullbackToFiber}.  Thus an irreducible module $V_\l\in
6608: \IrrRep(L_{P_\l})$ appears in the cohomology of
6609: \eqref{eqnMicroSupportOnFiberCalculation} if and only if
6610: $H(\i_P^*\ihat_Q^!\M)_V\neq 0$ for some $V=\tilde V_h\otimes V_\l\in
6611: \IrrRep(L_P)$.  Furthermore in this case $ Q_\l \in [Q_{V_\l}^{R,\l},
6612: Q_{V_\l}^{\prime R,\l}]$ if and only if $Q\in [Q_V, Q'_V]$.
6613: \end{proof}
6614: 
6615: \section{Functoriality under Inverse Images (continued)}
6616: \label{sectFunctorialityMicroSupportEqualRank}
6617: Let $F_R$ be a stratum of a Satake compactification $\Xstar_\sigma= \G\back
6618: \Dstar_\sigma$.  In this section we present a generalization of Proposition
6619: ~\ref{ssectWeakMicroSupportClosedInverseImage} (replacing $\mS_w$ by $\mS$)
6620: in the case that $\ihat_R$ is the inclusion of $\pi^{-1}(F_R)=
6621: X_R(L_{R,\l})$, under the assumption that $\Dstar_\sigma$ has equal-rank
6622: rational boundary components.  We begin with two lemmas, for which this
6623: assumption is not necessary.
6624: 
6625: We will use the notation of \S\S\ref{ssectPullbackToFiber} and
6626: \ref{sectSatakeCompactifications}.
6627: 
6628: \begin{lem}
6629: \label{ssectTransferLemma}
6630: Assume $V\preccurlyeq_\pm V'$ where $V$ and $V'$ are irreducible $L_P$- and
6631: $L_{P'}$-modules respectively.  Set $R=P^\dag$ and $R'={P'}^\dag$ and
6632: assume $P'\cap R = P$.  Then
6633: \begin{equation}
6634: R'\cap (Q_V^R\vee P') = Q_{V'}^{R'} \qquad \text{and} \qquad R'\cap
6635: (Q_V^{\prime R}\vee P') = Q_{V'}^{\prime R'} \label{eqnQTransfer}
6636: \end{equation}
6637: \end{lem}
6638: 
6639: \begin{proof}
6640: For $\al\in \D_P$ let $\atilde$ denote the unique element of $\D$ whose
6641: restriction to $\sa_P$ is $\al$.  In general
6642: $\langle\al,\al_0\spcheck\rangle\neq0$ for $\al$, $\al_0\in\D_P$ if and
6643: only if $\varPsi\cup\{\atilde,\atilde_0\}$ is connected for some
6644: $\varPsi\subseteq \D^P$.
6645: This is easy to verify if $\#\D^P=1$ and then one uses induction.
6646: 
6647: We claim that
6648: \begin{equation}
6649: R\vee P' \ge R'. \label{eqnTwoBoundarySets}
6650: \end{equation}
6651: To see this, note that if $\al\in \D_P$ and $\al\notin
6652: \D_P^R\cup\D_P^{P'}$, then $\{\tilde \al\}\cup \kap(\D^P)$ is
6653: $\sigma$-connected and $\al|_{\sa_{P'}}=\al'\neq 0$.  The first assertion
6654: implies that $\{\tilde \al\}\cup \kap(\D^{P'})$ is $\sigma$-connected.
6655: Thus $\al' = \tilde \al|_{\sa_{P'}} \notin \D_{P'}^{R'}$ and therefore
6656: $\al\notin\D_P^{R'}$.  This proves the claim.
6657: 
6658: Now let $\al'\in\D_{P'}^{R'}$ and let $\al$ be the unique element of
6659: $\D_P\setminus\D_P^{P'}$ for which $\al|_{\sa_{P'}}=\al'$.  Equation
6660: \eqref{eqnTwoBoundarySets} implies in fact that $\al\in 
6661: \D_P^R\setminus\D_P^{P'}$.  We claim that
6662: \begin{equation}
6663: \al|_{\sa_P^{P'}}=0\ .
6664: \label{eqnRootOrthogonality}
6665: \end{equation}
6666: To prove the claim let $\al_0\in \D_P^{P'}$.  As above if
6667: $\langle\al,\al_0\spcheck\rangle\neq 0$ then
6668: $\varPsi\cup\{\atilde,\atilde_0\}$ is connected for some $\varPsi\subseteq
6669: \D^P$.  But the assumption $P'\cap R = P$ implies that $\atilde_0\notin
6670: \o(\kap(\D^P))$ and hence $\atilde_0\in \kap(\D^{P'})$.  It follows that
6671: $\Psi\subseteq \kap(\D^{P'})$ and thus $\{\atilde\}\cup\kap(\D^{P'})$ is
6672: $\sigma$-connected.  But then $\al'=\tilde \al|_{\sa_{P'}}\notin
6673: \D_{P'}^{R'}$, a contradiction.  Hence $\langle\al,\al_0\spcheck\rangle =
6674: 0$ for all $\al_0\in\D_P^{P'}$, which proves \eqref{eqnRootOrthogonality}.
6675: 
6676: Now equation \eqref{eqnRootOrthogonality} implies that
6677: \begin{equation}
6678: \langle\xi_{V'} + \hsr,\al'{}\spcheck\rangle =
6679: \langle\xi_V+\hsr,\al\spcheck\rangle
6680: \label{eqnRootProjection}
6681: \end{equation}
6682: in view of  $\xi_{V'} + \hsr_{P'}=
6683: (\xi_V+\hsr)|_{\sa_{P'}}$.  It follows that $\al'\in\D_{P'}^{Q_{V'}^{R'}}$
6684: if and only if $\al\in \D_P^{Q_V^R}$, that is, $\al'\in \D_{P'}^{Q_V^R\vee
6685: P'}$.  Since  $\D_{P'}^{(Q_V^R\vee P')\cap R'} = \D_{P'}^{(Q_V^R\vee P')}
6686: \cap \D_{P'}^{R'}$, this proves the first equality of \eqref{eqnQTransfer};
6687: the other follows similarly.
6688: \end{proof}
6689: 
6690: \begin{lem}
6691: \label{ssectEqualRankTransferLemma}
6692: Assume $V\preccurlyeq_\pm V'$ where $V$ and $V'$ are irreducible $L_P$- and
6693: $L_{P'}$-modules respectively.  Set $R=P^\dag$ and $R'={P'}^\dag$ and
6694: assume $P'\cap R = P$.  If $V|_{M_P}$ is conjugate self-contragredient and
6695: $F_{R'}$ is equal-rank, then $V'|_{M_{P'}}$ is conjugate
6696: self-contragredient.
6697: \end{lem}
6698: 
6699: \begin{proof}
6700: Write $V' = \widetilde{V'_h}\otimes V'_\l$, a tensor
6701: product of irreducible modules according to the almost direct product
6702: decomposition $L_{P'}=\widetilde{L_{P',h}}L_{P',\l}$.  The module
6703: $\widetilde{V'_h}$ is automatically conjugate self-contragredient since
6704: $\widetilde{L_{P',h}} \cong \widetilde{L_{R',h}}$ has no $\QQ$-split
6705: center and $D_{R',h}$ is equal-rank.  To analyze $V'_\l$, note that
6706: $P'\cap R = P$ and $R=P^\dag$ implies that the parabolic $P/N_{P'}$ of
6707: $L_{P'}= \widetilde{L_{P',h}}L_{P',\l}$ is obtained by removing simple
6708: roots of $\widetilde{L_{P',h}}$.  It follows that $N_P^{P'}\subseteq
6709: \widetilde{L_{P',h}}$ and that $L_P$ factors as an almost direct
6710: product
6711: \begin{equation*}
6712: L_P = \widetilde{L_{P,h}}L_{P,\l}^{P',h} L_{P',\l}.
6713: \end{equation*}
6714: Let $\Res_{L_{P',\l}}^{L_P}$ denote the resulting functor that restricts a
6715: representation of $L_P$ to a representation of $L_{P',\l}$.  The
6716: preceding discussion shows that $\Res_{L_{P',\l}}^{L_P} H^\l(\n_P^{P'};V')$
6717: is isomorphic to a sum of copies of $V'_\l$.  On the other hand, $V$
6718: occurs as a component of $H^\l(\n_P^{P'}; V')$, so $\Res_{L_{P',\l}}^{L_P}
6719: V$ is also a sum of copies of $V'_\l$.  Since $V|_{M_P}$ is conjugate
6720: self-contragredient, this implies $V'_\l|_{M_{P',\l}}$ is conjugate
6721: self-contragredient.
6722: \end{proof}
6723: 
6724: \begin{prop}
6725: \label{ssectMicroSupportInverseImages}
6726: Let $F_R$ be a stratum of a Satake compactification $\Xstar_\sigma$ and
6727: assume that all rational boundary components $D_{R',h}\ge D_{R,h}$ are
6728: equal-rank.  Let $\ihat_R\colon  X_R(L_{R,\l}) = \pi^{-1}(F_R) \hookrightarrow
6729: \Xhat$ be the inclusion.  If $\M$ is an $\L_{\Xhat}$-module, then
6730: \begin{equation*}
6731: \mS(\ihat_R^*\M) \subseteq \ihat_R^*\mS(\M) \qquad
6732: \text{and}\qquad
6733: \mS(\ihat_R^!\M) \subseteq \ihat_R^!\mS(\M)\ .
6734: \end{equation*}
6735: Furthermore the estimates \eqref{eqnWeakMSInverseImageDegree} and
6736: \eqref{eqnWeakMSInverseImageProperSupportDegree} hold in the following
6737: strengthened form\textup: we can assume that $V\preccurlyeq_\pm \tilde V$
6738: may be factored as $V=V_0\preccurlyeq_\pm V_1 \preccurlyeq_\pm \dots
6739: \preccurlyeq_\pm V_N=\tilde V$ such that
6740: \begin{align}
6741: \label{eqnRemoveHermitianRoot}
6742: P_{i+1}\cap P_i^\dag = P_i &\qquad \text{for $0\le i < N$,} \\
6743: \label{eqnConjugateSelfContragredient}
6744: V_i|_{M_{P_i}} \text{ is conjugate self-contragredient} &\qquad \text{for
6745: $0\le i\le N$, and} \\
6746: \label{eqnDominantCone}
6747: (\xi_{V_i}+\hsr)|_{\sa_{P_i}^{P_{i+1}}}\in \pm \sa_{P_i}^{P_{i+1}*+} &\qquad
6748: \text{for $0\le i < N$.}
6749: \end{align}
6750: \end{prop}
6751: 
6752: \begin{proof}
6753: We first note that by Lemma ~\ref{ssectRationalBoundaryComponents} the
6754: $\sigma$-saturated parabolic $\QQ$-subgroups $R$ have $\QQ$-type $\o(\U)$
6755: for some $\sigma$-connected $\U\subseteq \D$.  Equivalently, $R$ is
6756: $\sigma$-saturated if and only if
6757: \begin{equation}
6758: \g\spcheck\not\perp \kap(\D^R)\cup\{\lsb\QQ \u\}\qquad \text{for all
6759: $\g\in\D\setminus\D^R$}\ ,
6760: \label{eqnSigmaSaturatedCondition}
6761: \end{equation}
6762: where $\lsb\QQ\u$ is the highest $\QQ$-weight of $\sigma$.
6763: 
6764: Once again we only consider the case of an irreducible $L_P$-module $V\in
6765: \mS(\ihat_R^*\M)$; the case $V\in \mS(\ihat_R^!\M)$ involves only minor
6766: changes.  Since $P\in\Pl(X_R(L_{R,\l}))$, we have $R=P^\dag$, a
6767: $\sigma$-saturated parabolic $\QQ$-subgroup.  Let $U(F_R)=
6768: \coprod_{F_{R'}\ge F_R} F_{R'}$ be the open star neighborhood of the
6769: stratum $F_R\subseteq \Xstar_\sigma$ and set $W=\pi^{-1}(U(F_R))$.  In
6770: order to apply the proof of Proposition
6771: ~\textup{\ref{ssectWeakMicroSupportClosedInverseImage}} to $V\in
6772: \mS(\ihat_R^*\M)$, we must replace $\M$ by its restriction to $W$; this is
6773: permissible by Proposition ~\ref{ssectMicroSupportOpenInverseImage}.  Now
6774: recall that at one point in that proof we obtain an irreducible
6775: $L_{P'}$-module $V'\in \mS_w(\ihat_{R'}^*\ihat_{S'}^!\M)$ satisfying
6776: $V\preccurlyeq_+ V'$ and $P'\cap R = P$ and we were free to choose $R'$
6777: provided that $R'\cap (Q\vee U) \le Q_{V'}^{\prime R'}$.  We now choose $R'
6778: = {P'}^\dag$.  Since $Q\in [Q_V^R,  Q_V^{\prime R}]$ and $R'\cap U=P'$ by
6779: \eqref{eqnTwoBoundarySets}, Lemma ~\ref{ssectTransferLemma} implies these
6780: conditions are satisfied and Lemma ~\ref{ssectEqualRankTransferLemma}
6781: implies that $V'|_{M_{P'}}$ is conjugate self-contragredient given that
6782: $V|_{M_P}$ is, so $V' \in \mS(\ihat_{R'}^*\ihat_{S'}^!\M)$.  Also
6783: $(\xi_V+\hsr)|_{\sa_P^{P'}} \in \sa_P^{P'*+}$ since by definition of $P'$
6784: all $\al\in\D_P^{P'}$ satisfy $\langle\xi_V+\hsr,\al\spcheck\rangle>0$.
6785: 
6786: Finally we need to verify that the passing by induction from $V$ to
6787: $V'$ preserves our new
6788: assumptions.  Observe that $R'<S'$ and thus $S'$ is also
6789: $\sigma$-saturated by the criterion
6790: \eqref{eqnSigmaSaturatedCondition}.  Let $U^{S'}(F_{R'})=U(F_{R'})\cap
6791: F_{S'}^*$ be the open star neighborhood of the stratum $F_{R'}$ in the
6792: induced Satake compactification $F_{S'}^*= \cl{F_{S'}}$.  Set $W' =
6793: \pi^{-1}(U^{S'}(F_{R'}))$.  Consider the diagram of inclusions where
6794: the vertical arrows are open embeddings:
6795: \begin{equation*}
6796: \xymatrix{
6797: {\Xhat_{R'}\cap W} \ar@{^{ (}->}[r]^{\ihat_{R'}} & {\Xhat_{S'}\cap W} \\
6798: {X_{R'}(L_{R',\l})}  \ar@{^{ (}->}[r]_{\ihat_{R'}}
6799: \ar@{^{ (}->}[u]^{k_{R'}}
6800: & {W'\rlap{\ .}} \ar@{^{ (}->}[u]_{k_{S'}}
6801: }
6802: \end{equation*}
6803: By Proposition ~\ref{ssectMicroSupportOpenInverseImage},
6804: $V'\in\mS(k_{R'}^* \ihat_{R'}^*\ihat_{S'}^!\M) =
6805: \mS(\ihat_{R'}^*k_{S'}^*\ihat_{S'}^!\M)$, and we can use induction.
6806: \end{proof}
6807: 
6808: \section{The Basic Lemma}
6809: \label{sectBasicLemma}
6810: We now address the issue of estimating the term $[\tilde V:V]=\l(w)$ in
6811: \eqref{eqnWeakMSInverseImageDegree} and
6812: \eqref{eqnWeakMSInverseImageProperSupportDegree} within the context of
6813: Proposition \ref{ssectMicroSupportInverseImages}.  Our tool is a basic
6814: lemma which, under the assumption that $V|_{M_P}$ is conjugate
6815: self-contragredient, establishes the fundamental relationship between the
6816: geometry of $\xi_V$ {\itshape vis a vis\/} an $S_P^{\tilde P}$-root $\al$
6817: and the contribution to the length of $w$ attributable to $\al$.  This
6818: basic lemma is important in other contexts; a simple form already occurred
6819: as Lemma~\ref{ssectSimpleBasicLemma}.
6820: 
6821: For simplicity we will assume that $\tilde V=E$ is an irreducible
6822: $G$-module.  We use the notation of \S\ref{ssectSimpleBasicLemma}.
6823: 
6824: \subsection{}
6825: \label{ssectRealSubmodules}
6826: Let $\h$ be a fundamental Cartan subalgebra of $\levi_P$ and let
6827: $\Phi^+(\levi_{P\CC},\h_\CC)$ be a $\theta$-stable positive system for
6828: $\levi_P$.  Let $V$ be an irreducible $L_P$-module with highest weight
6829: $\u\in\h_\CC^*$ and assume $V|_{M_P}$ is conjugate self-contragredient.  We
6830: have defined a reductive $\RR$-subgroup $L_P(\u)\subseteq L_P$ in
6831: \S\S\ref{ssectWeightCentralizer} and \ref{ssectVcentralizer} with roots
6832: \begin{equation}
6833: \{\,\g\in\Phi(\levi_{P\CC},\h_{\CC}) \mid
6834: \langle\u,\g\spcheck\rangle=0\,\}\ .
6835: \end{equation}
6836: 
6837: Via a lift of $L_P$ to $P$, we obtain an adjoint representation of
6838: $\levi_P$ (and hence $\levi_P(\u)$) on $\n_{P\CC}$.  There is a unique
6839: decomposition
6840: \begin{equation*}
6841: \n_{P\CC} = \bigoplus_{\substack{\text{ $\levi_P(\u)$-irreducible}\\
6842:                          \text{submodules} \\
6843:                          F\subseteq \n_{P\CC}}}       F,
6844: \end{equation*}
6845: which coarsens the decomposition into root spaces.  Set $\t'_P = -\t_P$ and
6846: observe that if $F$ is an irreducible component as above with weights
6847: $\Phi(F,\h_\CC)$, then $\t'_P(\Phi(F,\h_\CC))$ is the set of weights of
6848: another irreducible component, which we denote $\t'_P(F)$; this follows
6849: from \eqref{eqnTauInvariant}.  In this way we obtain an action of $\t'_P$
6850: on the $\levi_P(\u)$-irreducible components of $\n_{P\CC}$,
6851: generalizing the action of $\t'_P$ on the roots of $\n_{P\CC}$.  Let%
6852: %
6853: \footnote{The quantities $\levi_P(\u)$ and $\n_P(\u)$ are
6854: defined quite differently despite the similarity of notation.  We trust
6855: that with this warning there should be no confusion.}
6856: %
6857: \begin{equation*}
6858: \n_P(\u) = \bigoplus_{\t'_P(F)=F} F ;
6859: \end{equation*}
6860: since the action of $\levi_P(\u)$ on $\n_{P\CC}$ preserves $\n_{\al\CC}$
6861: for any $\al\in\Phi(\n_P,\sa_P)$, we may likewise define $\n_\al(\u)$.
6862: Although $\n_P(\u)$ and $\n_\al(\u)$ depend on the choice of a lift of
6863: $L_P$ to $G$, their dimensions are independent of this choice.  The
6864: dimension of $\n_P(\u)$ can vary depending on the choice of
6865: $\Phi^+(\levi_{P\CC},\h_\CC)$ however; let $\n_P(V)$ denote any one of the
6866: $\n_P(\u)$ with maximal dimension.
6867: 
6868: \begin{BasicLemma}
6869: \label{ssectBasicLemma}
6870: Let $P$ be a parabolic $\QQ$-subgroup and let $w\in W_P$.  Let
6871: $V=H^{\l(w)}(\n_P;E)_w$ have highest weight $\u$ with respect to a
6872: $\theta$-stable positive system.  Assume $V|_{M_P}$ is conjugate
6873: self-contragredient.  For any $\al\in\Phi(\n_P,\sa_P)$ we have\textup:
6874: \begin{enumerate} 
6875: \item $\langle\xi_V+\hsr,\al\spcheck\rangle \le 
6876: 0 \implies 
6877: \l_\al(w) \ge \frac12(\dim \n_\al+\dim \n_\al(\u))$.
6878: \label{itemBasicLemmaLessThan}
6879: \item 
6880: $\langle\xi_V+\hsr,\al\spcheck\rangle = 0 \implies \l_\al(w) = \frac12\dim
6881: \n_\al$ and $\n_\al(\u)=0$.
6882: \label{itemBasicLemmaEqual}
6883: \item $\langle\xi_V+\hsr,\al\spcheck\rangle\ge 0 \implies \l_\al(w) \le 
6884: \frac12(\dim \n_\al-\dim \n_\al(\u))$.
6885: \label{itemBasicLemmaGreaterThan}
6886: \end{enumerate}
6887: Furthermore if $F\subseteq \n_{\al\CC}$ is an irreducible
6888: $\levi_P(\u)$-submodule, then $\Phi_w$ contains respectively
6889: \itemref{itemBasicLemmaLessThan} at least one,
6890: \itemref{itemBasicLemmaEqual} exactly one,
6891: \itemref{itemBasicLemmaGreaterThan} at most one of
6892: $\Phi(F,\h_\CC)$ and $\Phi(\t'_P(F),\h_\CC)$.
6893: \end{BasicLemma}
6894: 
6895: \begin{proof}
6896: Let $W^{L_P(\u)}$ denote the Weyl group of $L_P(\u)$ and let
6897: $\lambda\in\h_\CC$ be the highest weight of $E$ relative to $\Phi^+ =
6898: \Phi^+(\levi_{P\CC},\h_\CC) \cup \Phi(\n_{P\CC},\h_\CC)$.  We begin by
6899: establishing
6900: \begin{equation}
6901: s\in W^{L_P(\u)} \quad \Rightarrow \quad s\Phi_w =\Phi_w \text{ and }
6902: sw\lambda = w\lambda \label{eqnCartier}
6903: \end{equation}
6904: following \cite{refnCartier}.  Express the highest weight $\u = w(\lambda +
6905: \hsr)-\hsr$ of $V$ as $w\lambda -{\sum_{\g\in\Phi_w}\g}$.  Since $s$ may be
6906: written as a product of reflections in simple coroots orthogonal to $\u$ we
6907: see that $\u=s\u$ and hence
6908: \begin{equation*}
6909: w\lambda -{\sum_{\g\in\Phi_w}\g} =\u = s\u =s w\lambda -
6910: {\sum_{\g\in s\Phi_w} \g}.
6911: \end{equation*}
6912: Note that $s\Phi_w\subseteq \Phi^+$; by canceling the terms of the right
6913: hand sum that lie in $\Phi_w$ with the corresponding terms of the left hand
6914: sum one may compute that
6915: \begin{equation}
6916: \lambda = w^{-1}sw\lambda - {\sum_{\g\in (w^{-1}\Psi)\cap \Phi^+} \g},
6917: \label{eqnLambdaFormula}
6918: \end{equation}
6919: where $\Psi = s\Phi_w\cup -(\Phi^+\setminus s\Phi_w)$.  Since $\lambda$ is
6920: dominant, we may replace $w^{-1}sw\lambda$ in \eqref{eqnLambdaFormula} by
6921: $\lambda - \sum_i m_i
6922: \al_i$ for $\al_i$ simple and $m_i\ge 0$.  In addition since the
6923: roots of $(w^{-1}\Psi)\cap \Phi^+$ being subtracted in \eqref{eqnLambdaFormula}
6924: are positive, it follows that all $m_i=0$ and that
6925: $(w^{-1}\Psi)\cap \Phi^+ = \emptyset$; these two facts are equivalent to
6926: the desired \eqref{eqnCartier}.
6927: 
6928: Now consider an $\levi_P(\u)$-irreducible submodule $F\subseteq
6929: \n_{\al\CC}$.  Since by \eqref{eqnCartier}
6930: $\Phi_w$ is stable under $W^{L_P(\u)}$, either all extremal weights of
6931: $\Phi(F,\h_\CC)$ belong to $\Phi_w$ or none do.  Since furthermore membership
6932: in $\Phi_w$ is given by a linear inequality on positive
6933: roots and $\Phi(F,\h_\CC)$ is the convex hull of the extremal weights, this
6934: implies that either all of
6935: $\Phi(F,\h_\CC)$ is contained in $\Phi_w$ or none of it is.
6936: An application of Lemma ~\ref{ssectSimpleBasicLemma} to any $\g\in
6937: \Phi(F,\h_\CC)$ concludes the proof.
6938: \end{proof}
6939: 
6940: \begin{cor}
6941: \label{ssectBasicLemmaCor}
6942: Let $P$ be a parabolic $\QQ$-subgroup and let $V$ be an irreducible
6943: constituent of $H^i(\n_P;E)$ for which $V|_{M_P}$ is conjugate
6944: self-contragredient.  Then
6945: \begin{align*}
6946: (\xi_V+\hsr)|_{\sa_P} \in -\sa^{*+}_P &\implies i \ge \frac12(\dim
6947: \n_P+\dim\n_P(V)) \\
6948: \intertext{and}
6949: (\xi_V+\hsr)|_{\sa_P} \in \sa^{*+}_P &\implies i \le \frac12(\dim
6950: \n_P-\dim\n_P(V)). 
6951: \end{align*}
6952: \end{cor}
6953: 
6954: \begin{rem}
6955: All results in this section and their proofs generalize immediately to the
6956: situation in which $E$ is replaced by a ``virtual'' irreducible $G$-module,
6957: whose highest weight $\lambda$ is allowed to be any dominant element of
6958: $\h_\CC^*$, not necessarily integral.  They also generalize to parabolic
6959: $k$-subgroups where $k$ is any subfield of $\RR$.
6960: \end{rem}
6961: 
6962: \section{The Basic Lemma in the Presence of Equal-Rank Real Boundary
6963: Components}
6964: \label{sectEqualRankBasicLemma}
6965: We need to replace the term $\dim \n_P(V)$ appearing in Corollary
6966: ~\ref{ssectBasicLemmaCor} by a more geometric expression.  This can be
6967: done in the presence of a Satake compactification with equal-rank
6968: {\itshape real\/} boundary components.
6969: 
6970: \subsection{}
6971: \label{ssectVcentralizerRationalFactors}
6972: Let $V$ be an irreducible $L_P$-module and assume that $V|_{M_P}$ is
6973: conjugate self-contragredient.  Let $D_P(V)$ be defined as in
6974: \S\ref{ssectVcentralizer}.  Let $D\subseteq \lsb\RR\Dstar_\sigma$ be a
6975: Satake compactification.  The decomposition \eqref{eqnLevisRationalFactors}
6976: induces a decomposition of symmetric spaces
6977: \begin{equation}
6978: D_P(V)= D_{P,h}(V)\times D_{P,\l}(V)\ .
6979: \end{equation}
6980: Here $D_{P,h}(V)$ is defined using $\Res_{\widetilde{L_{P,h}}}^{L_P} V$
6981: and similarly for $D_{P,\l}(V)$; see the last comment in
6982: \S\ref{ssectVcentralizer}.
6983: 
6984: \begin{lem}
6985: \label{ssectEqualRankLemma}
6986: Let $P$ be a parabolic $\QQ$-subgroup and let $D\subseteq \lsb\RR
6987: \Dstar_\sigma$ be a Satake compactification for which we assume all
6988: real boundary components $D_{R',h}\ge D_{P^\dag,h}$ are equal-rank.
6989: Let $V$ be an irreducible constituent of $H(\n_P;E)$ for which
6990: $V|_{M_P}$ is conjugate self-contragredient.  Then
6991: \begin{equation*}
6992: \dim \n_P(V) \ge \dim \sa_P^G + \dim D_{P,\l}(V) \ .
6993: \end{equation*}
6994: \end{lem}
6995: \begin{rem*}
6996: A similar result was proved by Borel in
6997: \cite[\S5.5]{refnBorelVanishingTheorem} under the weaker assumptions that
6998: $D$ and $D_{P^\dag,h}$ are equal-rank and a certain condition (B) holds.
6999: The condition (B) needed to be verified for each desired case and Borel
7000: expressed the hope that it could be replaced by a more conceptual argument.
7001: The proof offered below, while not transparent, is at least free from case
7002: by case analysis.
7003: 
7004: Note also that condition (B) does not hold%
7005: %
7006: \footnote{In \cite[\S6.6]{refnBorelVanishingTheorem} it is mistakenly
7007: asserted that condition (B) holds in this case.}
7008: %
7009: in one case of interest where the lemma above {\itshape does\/} apply:
7010: $G=\SO(p,q)$ with $p+q$ odd with the Satake compactification in which
7011: $\sigma$ is connected only to the short simple $\RR$-root.  In this case
7012: there are imaginary noncompact roots in $\levi_P$ (relative to a
7013: fundamental Cartan subalgebra) which are strongly orthogonal to all real
7014: roots in $\n_P$; the hypothesis that the intermediate real boundary
7015: components are equal-rank and its use below in Sublemma
7016: ~\ref{ssectEqualRankSubLemma} was needed precisely to handle this
7017: situation.
7018: \end{rem*}
7019: 
7020: \begin{proof}[Proof of Lemma ~\textup{\ref{ssectEqualRankLemma}}]
7021: We first establish notation.  Let $\theta$ be a Cartan involution of
7022: $L_P(\RR)$ with Cartan decomposition $\levi_P = \k_P + \p_P + \sa_P$.  Let
7023: $\h= \hb_{P,\k}+\hb_{P,\p} +\sa_P$ be a fundamental $\theta$-stable Cartan
7024: subalgebra of $\levi_P$ and choose a $\theta$-stable positive system
7025: $\Phi^+(\levi_{P\CC},\h_\CC)$ for $\levi_{P\CC}$; thus $\t_P=\theta$.  Let
7026: $\u$ be the highest weight of $V$ and let $L_P(\u)$ be defined as in
7027: \S\ref{ssectWeightCentralizer} with Cartan decomposition $\k_P(\u) +
7028: \p_P(\u) +\sa_P$.  Let $\Phi^+(\p_P(\u)_\CC,\hb_{P,\k\CC})$ denote the
7029: positive nonzero weights of $\hb_{P,\k\CC}$ in $\p_P(\u)_\CC$, counted with
7030: multiplicity.  We extend the preceding notation to $L_{P,h}$ and $L_{P,\l}$
7031: by adding the appropriate subscript.  Finally note that any reductive
7032: subalgebra $\r\subseteq \levi_{P\CC}$ acts on $\n_{P\CC}$ by the adjoint
7033: action of a lift of $\levi_{P\CC}$ and define $\n_P(\r)$ to be the sum of
7034: the irreducible $\r$-submodules whose set of weights are $\t'_P$-stable;
7035: thus $\n_P(\levi_P(\u))=\n_P(\u)$ as in \S\ref{ssectRealSubmodules}.  It
7036: will suffice to prove $\dim \n_P(\u) \ge \dim \sa_P^G + \dim D_{P,\l}(\u)$.
7037: 
7038: Clearly
7039: \begin{equation}
7040: \label{eqnLinearDimension}
7041: \dim \sa_P^G + \dim D_{P,\l}(\u) =
7042: 2\cdot\#\Phi^+(\p_{P,\l}(\u)_\CC,\hb_{P,\k\CC}) + \dim \hb_{P,\p} + \dim
7043: \sa_P^G \ .
7044: \end{equation}
7045: We begin by bounding the last two terms.  Consider $\n_P(\h_\CC)$ as
7046: defined above; the set $\Phi(\n_P(\h_\CC),\h_\CC)$ consists of the positive
7047: roots $\g$ satisfying $\t'_P \g = \g$, that is, the real roots.  We claim
7048: that
7049: \begin{equation}
7050: \label{eqnRealRoots}
7051: \dim \n_P(\h_\CC) \ge \dim \hb_{P,\p} + \dim \sa_P^G\ .
7052: \end{equation}
7053: In fact there exists a set of strongly orthogonal roots
7054: $\{\d_1,\dots,\d_r\}$ satisfying $\t'_P \d_i = \d_i$ and which span
7055: $\hb_{P,\p}^*+\sa_P^G{}^*$.  To see this, proceed as in
7056: \cite{refnBorelVanishingTheorem} to write the centralizer in
7057: $\lsp0\mathfrak g$ of $\hb_{P,\k}$ as $\hb_{P,\k} + \mathfrak z_1$.  Then
7058: $\mathfrak z_1$ has both a split Cartan subalgebra $\hb_{P,\p}+\sa_P^G$
7059: and, since by hypothesis $\lsp0G =L_{G,h}$ is equal-rank, a compact Cartan
7060: subalgebra.  The existence of $\{\d_1,\dots,\d_r\}$ then follows from
7061: \cite{refnKostantConjugacyCartan}, \cite[Prop.~ 11]{refnSugiura},
7062: \cite{refnSugiuraCorrection}.
7063: 
7064: We now attempt to bound the rest of \eqref{eqnLinearDimension}.
7065: For $\g_\k\in \Phi^+(\p_{P,\l}(\u)_\CC,\hb_{P,\l,\k\CC})$ choose
7066: $\g\in\Phi^+(\levi_{P,\l}(\u)_\CC,\hb_{P,\l\CC})$ such that
7067: $\g|_{\hb_{P,\l,\k\CC}} = \g_\k$.  The root $\t_P\g$ is also in
7068: $\Phi^+(\levi_{P,\l}(\u)_\CC,\hb_{P,\l\CC})$ since
7069: $\t_P(\u|_{\hb_P})=\u|_{\hb_P}$.  Let $\r_\g$ denote the reductive
7070: subalgebra $\h_\CC\subseteq \r_\g \subseteq \levi_P(\u)_\CC$ with root
7071: system $\Phi\cap \Span\{\g,\t_P\g\}$.  Clearly
7072: \begin{equation*}
7073: \Phi(\n_P(\h_\CC),\h_\CC) \subseteq \Phi(\n_P(\r_\g),\h_\CC) \subseteq
7074: \Phi(\n_P(\levi_P(\u)_\CC),\h_\CC)= \Phi(\n_P(\u),\h_\CC).
7075: \end{equation*}
7076: We will see below in Sublemma ~\ref{ssectEqualRankSubLemma}%
7077: \itemref{itemEqualRankTauNoninvariant}\itemref{itemEqualRankTauInvariant}
7078: that there exists a pair of roots $\eta\neq\t'_P\eta \in
7079: \Phi(\n_P(\r_\g),\h_\CC) \setminus \Phi(\n_P(\h_\CC),\h_\CC) \subseteq
7080: \Phi(\n_P(\u),\h_\CC)$.  This would imply
7081: \begin{equation}
7082: \dim \n_P(\u) - \dim \n_P(\h_\CC) \ge
7083: 2\cdot\#\Phi^+(\p_{P,\l}(\u)_\CC,\hb_{P,\k\CC})
7084: \end{equation}
7085: (and hence prove the lemma in view of \eqref{eqnLinearDimension} and
7086: \eqref{eqnRealRoots}) except that we have to deal with the possibility that
7087: the sets $\Phi(\n_P(\r_\g),\h_\CC) \setminus \Phi(\n_P(\h_\CC),\h_\CC)$ for
7088: the different $\g_\k$ may not be disjoint.
7089: 
7090: Since $|\g|=|\t_P\g|$ and no multiple of $\g-\t_P\g$ can be a root (recall
7091: that $\levi_{P\CC}$ has no roots negated by $\t_P$), one easily checks that
7092: $\{\g,\t_P\g\}$ is a base for the root system of $\r_\g$ and the type is
7093: $A_1$ (if $\g=\t_P\g$), $A_1\times A_1$, or $A_2$.  Now consider $\g_\k\neq
7094: \g_\k'\in \Phi^+(\p_{P,\l}(\u)_\CC,\hb_{P,\l,\k\CC})$ and choose
7095: corresponding roots $\g\neq \g'
7096: \in\Phi^+(\levi_{P,\l}(\u)_\CC,\hb_{P,\l\CC})$.  Suppose that
7097: $\Phi(\n_P(\r_\g),\h_\CC)\setminus \Phi(\n_P(\h_\CC),\h_\CC)$ and
7098: $\Phi(\n_P(\r_{\g'}),\h_\CC)\setminus \Phi(\n_P(\h_\CC),\h_\CC)$ have
7099: nonempty intersection.  Then there exists a root $\eta\in
7100: \Phi(\n_{P\CC},\h_\CC)$ for which $\t'_P\eta-\eta$ is a nonzero
7101: $\ZZ$-linear combination of both $\{\g,\t_P\g\}$ and $\{\g',\t_P\g'\}$.
7102: Since $\t'_P\eta-\eta$ is $\t_P$-invariant, this means that $\g+\t_P\g$ and
7103: $\g'+\t_P\g'$ are scalar multiples of each other.  But $(1/2)(\g+\t_P\g)$
7104: belongs to either the reduced root system $\Phi(\levi_{P\CC},\h_\CC)$ (if
7105: $\g=\t_P\g$) or the reduced root system $\Phi(\k_{P\CC},\hb_{P,\k\CC})$ (if
7106: $\g\neq \t_P\g$), and similarly for $(1/2)(\g'+\t_P\g')$.  Thus $\g+\t_P\g$
7107: and $\g'+\t_P\g'$ can be scalar multiples of each other only if $\g$ (say)
7108: satisfies $\g\neq \t_P\g$ and $\g'$ satisfies $\g'=\t_P\g'$.  It follows
7109: from our comments earlier then that $\g'=\g+\t_P\g$ and $\r_\g$ has type
7110: $A_2$.
7111: 
7112: So $\Phi(\n_P(\r_\g),\h_\CC)\setminus \Phi(\n_P(\h_\CC),\h_\CC)$ can
7113: intersect with at most one other such set and it will suffice to show that
7114: in this case $\Phi(\n_P(\r_\g),\h_\CC)\setminus \Phi(\n_P(\h_\CC),\h_\CC)$
7115: contains an additional pair of roots $\eta'\neq \t'_P\eta'$.  But if such a
7116: pair of roots did not exist, the only nontrivial $\r_\g$-submodule of
7117: $\n_P(\r_\g)$ would have weights (after interchanging $\eta$ and
7118: $\t'_P\eta$ if necessary) $\{\eta, \eta+\g=\t'_P\eta-\t_P\g, \t'_P\eta\}$
7119: and hence $\{\eta,\g,\t_P\g\}$ would be the base of a root subsystem of
7120: type $A_3$.
7121: Sublemma~\ref{ssectEqualRankSubLemma}\itemref{itemEqualRankAThree}
7122: following shows this situation cannot occur.
7123: \end{proof}
7124: 
7125: \begin{sublem}
7126: \label{ssectEqualRankSubLemma}
7127: As in the proof of Lemma ~\textup{\ref{ssectEqualRankLemma}} let
7128: $\g_\k\in \Phi^+(\p_{P,\l\CC},\hb_{P,\l,\k\CC})$ and choose $\g\in
7129: \Phi^+(\levi_{P,\l\CC},\hb_{P,\l\CC})$ such that
7130: $\g|_{\hb_{P,\l,\k\CC}}=\g_\k$.
7131: \begin{enumerate}
7132: \item If $\t_P\g \neq \g$, then there exists a root $\d=\t'_P\d
7133: \in \Phi(\n_{P\CC},\h_\CC)$ such that {\rm(}after interchanging $\g$ and
7134: $\t_P\g$ if necessary\/{\rm)} $\eta= \d-\g$ and
7135: $\t'_P\eta=\d+\t_P\g$ are also roots.
7136: \label{itemEqualRankTauNoninvariant}
7137: \item If in \itemref{itemEqualRankTauNoninvariant} the roots
7138: $\{\eta,\g,\t_P\g\}$ are a base of a root subsystem of type $A_3$, then
7139: there also exists a root $\dtilde\neq \d$ with $\dtilde=\t'_P\dtilde$ and
7140: one of $\dtilde\pm \g$ a root.
7141: \label{itemEqualRankAThree}
7142: \item If $\t_P\g=\g$, then there exists a root $\eta \in
7143: \Phi(\n_{P\CC},\h_\CC)$ such that $\t'_P\eta=\eta+k\g$ for some
7144: $k\in\ZZ\setminus\{0\}$.
7145: \label{itemEqualRankTauInvariant}
7146: \end{enumerate}
7147: \end{sublem}
7148: \begin{proof}
7149: Let $\{\d_1,\dots,\d_r\}$ be a set of strongly orthogonal roots spanning
7150: $\hb_{P,\p}^*+\sa_P^{G\,*}$ and satisfying $\t'_P \d_i = \d_i$ as in the
7151: proof of Lemma ~\ref{ssectEqualRankLemma}.  If $\t_P\g\neq \g$ then
7152: $\g|_{\hb_{P,\p}+\sa_P^G}\neq 0$ and so there exists $\d=\d_i$ such that
7153: $\langle\g,\d\spcheck\rangle\neq 0$.  By interchanging $\g$ and $\t_P\g$ if
7154: necessary, we may assume $\langle\g,\d\spcheck\rangle >0$.  Then $\eta = \d
7155: -\g$ is a root which proves \itemref{itemEqualRankTauNoninvariant}.  Part
7156: \itemref{itemEqualRankAThree} will follow if we show there exists
7157: $\dtilde=\d_j\neq \d_i$ with $\langle\g,\dtilde\spcheck\rangle\neq 0$ as
7158: well.  But if such a $\dtilde$ did not exist, $\g|_{\hb_{P,\p}+\sa_P^G} = c
7159: \d$ for some scalar $c$ and it is easy to check that the conditions on $c$
7160: implied by the equalities $|\d-\g|=|\g|=|\d|$ and $2(\g,\t_P\g)/|\g|^2 =
7161: -1$ for a Weyl group invariant inner product on a root system of type $A_3$
7162: are inconsistent.
7163: 
7164: If $\t_P\g = \g$ we claim that there is a nontrivial $\g$-root string
7165: through some root in $\RR\g +\hb_{P,\p}^*+\sa_P^{G\,*}$.  On such a root
7166: string the operator $\t'_P$ acts as reflection in $\g$ and hence the root
7167: string is $\t'_P$-stable.  Thus part ~\itemref{itemEqualRankTauInvariant}
7168: follows from the claim if we set $\eta$ to be any root in the root string
7169: other than at the center.
7170: 
7171: The claim is obvious if $\d_i+\g$ or $\d_i-\g$ is a root for some $i$ we
7172: are done, so we assume that $\g$ is strongly orthogonal to
7173: $\{\d_1,\dots,\d_r\}$.
7174: Recall \cite[Chapter~ VII, ~\S7]{refnKnapp} that a parabolic $\RR$-subgroup
7175: $R$ is called {\itshape cuspidal\/}%
7176: %
7177: \footnote{This usage is  different from that in
7178: \cite{refnLanglandsFE} or \cite{refnBorelCasselman}; $R$ being cuspidal
7179: in those references corresponds here to $R$ being defined over $\QQ$.}
7180: %
7181: if $L_R/Z(L_R)$ is equal-rank, where $Z(L_R)$ denotes the center.  Let
7182: $R\subseteq P$ be a cuspidal parabolic $\RR$-subgroup with $\lsb\RR\hb_{R}
7183: = \hb_{P,\k}$ and split component $\lsb\RR \sa_{R} = \hb_{P,\p}+\sa_P$; to
7184: construct $R$ apply \cite[Prop.~7.87]{refnKnapp}.  Since
7185: $\t_{R}=\t_P=\theta$ on $\h$, $\g$ remains an imaginary noncompact root for
7186: $\levi_{R\CC}$.  Thus there is a Cayley transformation \cite[Chapter~ VI,
7187: ~\S7]{refnKnapp} which transforms $\g$ into a real root.  That is, there
7188: exists a inner automorphism $\mathbf c_{\g}$ of $\levi_{R\CC}$ such that
7189: $\h'\equiv \levi_R \cap \mathbf c_{\g}(\h_\CC)$ is a $\theta$-stable Cartan
7190: subalgebra and $\mathbf c_{\g}\circ(s_{\g}\circ\theta) = \theta \circ
7191: \mathbf c_{\g}$ on $\h$.  Let $R'\subset R$ be a cuspidal parabolic
7192: $\RR$-subgroup with $\lsb\RR\hb_{R'}= \mathbf
7193: c_{\g}(\ker\g\cap\lsb\RR\hb_R)$ and $\lsb\RR\sa_{R'}=\RR \mathbf
7194: c_\g(\g\spcheck)+ \lsb\RR\sa_R$.  The roots $\{\d_1,\dots,\d_r,\mathbf
7195: c_\g(\g)\}$ form a strongly orthogonal basis of $\lsb\RR\sa_{R'}^{G\,*}$
7196: and to prove our claim we need to demonstrate that there is a nontrivial
7197: $\mathbf c_\g(\g)$-root string through some root in
7198: $\lsb\RR\sa_{R'}^{G\,*}$.
7199: 
7200: From our hypotheses on the Satake compactification we now construct another
7201: independent set $\{\d'_1,\dots,\d'_{r+1}\}$ of roots spanning
7202: $\lsb\RR\sa_{R'}^{G\,*}$.  Define a sequence of parabolic $\RR$-subgroups
7203: \begin{equation*}
7204: R'=R'_0\subset R'_1 \subset \dots \subset R'_{r+1} = G
7205: \end{equation*}
7206: inductively by choosing $R'_i$ to have type $\DRR_{R'}^{R'_i}=\{\al_1,\dots
7207: ,\al_i\}$ where $\al_i\in\DRR_{R'}^{R',R_{i-1}^{\prime\dag}}$.  Thus the
7208: corresponding real Satake boundary component strictly increases at each stage,
7209: \begin{equation*}
7210: D_{R^{\prime\dag},h}= D_{R_0^{\prime\dag},h} \lneq  D_{R_1^{\prime\dag},h}
7211: \lneq \dots \lneq D_{R_{r+1}^{\prime\dag},h} = D \ .
7212: \end{equation*}
7213: On the other hand, since $L_{P,h}$ is equal-rank and $\g$ was a root of
7214: $\levi_{P,\l}$, we have $L_{R',h} = L_{P,h}$ and hence
7215: $D_{P^\dag,h}=D_{R^{\prime\dag},h}$.  We now claim that each parabolic
7216: subgroup $R'_i$ is cuspidal.  This follows since $L_{R'_i,\l}/Z(L_{R'_i})$
7217: is an almost direct factor of $L_{R'_{i-1},\l}/Z(L_{R'_{i-1}})$, which is
7218: equal-rank by induction, and $L_{R'_i,h}$ is equal-rank since the boundary
7219: component $D_{R'_i{}^\dag,h}$ is equal-rank by hypothesis.  Thus for each
7220: $i=1,\dots,r+1$ there exists a root $\d'_i$ which spans
7221: $\lsb\RR\sa_{R'_{i-1}}^{R'_i\,*}$ \cite{refnKostantConjugacyCartan},
7222: \cite{refnSugiura}, \cite{refnSugiuraCorrection}.
7223: 
7224: Since $\mathbf c_\g(\g)\in \lsb\RR\sa_{R'}^{G\,*}$ is nonzero there exists
7225: $1\le i \le r+1$ such that $\langle\mathbf
7226: c_\g(\g),\d'_i{}\spcheck\rangle\neq 0$.  The desired nontrivial $\mathbf
7227: c_\g(\g)$-root string would now exist provided $\mathbf c_\g(\g)\neq \pm
7228: \d'_i$.  To show $\mathbf c_\g(\g)$ cannot equal $\pm\d'_i$, write
7229: $\DRR_{R'}^R=\{\al_k\}$ where $1\le k \le r+1$.  Thus $\mathbf c_\g(\g)$ is
7230: a multiple of $\al_k$.  On the other hand, $\d'_i$ is proportional to the
7231: projection of $\al_i$ to $\lsb\RR\sa_{R_{i-1}'}^{R_i'}$, that is,
7232: $\d'_i\sim\al_i + \sum_{j < i} c_j \al_j$ where all $c_j=0$ if and only if
7233: $\al_i\in \lsb\RR\sa_{R_{i-1}'}^{R_i'}$.  So if $\mathbf c_\g(\g)$ were to
7234: equal $\pm \d'_i$ then $k=i$ and $\al_i\in \lsb\RR\sa_{R_{i-1}'}^{R_i'}$.
7235: This implies that the simple $\RR$-roots underlying $\al_i$ as opposed to
7236: $\{\al_1,\dots,\al_{i-1}\}$ belong to different components of $L_{R'_i,h}$.
7237: This implies that the condition $\al_i\in\DRR_{R'}^{R',R_{i-1}'{}^\dag}$
7238: actually means $\al_i\in\DRR_{R'}^{R',R'{}^\dag}$.  But since $\g$ is a root
7239: of $\levi_{P,\l}$, $\mathbf c_\g(\g)$ is a root of $\levi_{R,\l}$ and hence
7240: a linear combination of roots in $\DRR_{R'}^{R{}^\dag}=
7241: \DRR_{R'}^{R'{}^\dag}$ and so can't be proportional to $\al_i$.
7242: \end{proof}
7243: 
7244: \begin{cor}
7245: \label{ssectEqualRankBasicLemmaCor}
7246: Let $P$ be a parabolic $\QQ$-subgroup and let $D\subseteq \lsb\RR
7247: \Dstar_\sigma$ be a Satake compactification for which we assume all
7248: real boundary components $D_{R',h}\ge D_{P^\dag,h}$ are equal-rank.
7249: Let $V$ be an irreducible constituent of $H^i(\n_P;E)$ for which
7250: $V|_{M_P}$ is conjugate self-contragredient.  Then
7251: \begin{align*}
7252: (\xi_V+\hsr)|_{\sa_P} \in -\sa^{*+}_P &\implies i \ge \frac12(\dim
7253: \n_P+\dim \sa_P^G + \dim D_{P,\l}(V)) \\ \intertext{and}
7254: (\xi_V+\hsr)|_{\sa_P} \in \sa^{*+}_P &\implies i \le \frac12(\dim
7255: \n_P-\dim \sa_P^G - \dim D_{P,\l}(V)).
7256: \end{align*}
7257: \end{cor}
7258: \begin{proof}
7259: Combine Lemma ~\ref{ssectEqualRankLemma} and Corollary
7260: ~\ref{ssectBasicLemmaCor}.
7261: \end{proof}
7262: \begin{rem*}
7263: For the Baily-Borel-Satake compactification in the Hermitian case, the
7264: second inequality (for $\xi_V$ in a larger cone) is the content of
7265: \cite[Prop.~ 11.1]{refnSaperSternTwo}.  Shortly after
7266: \cite{refnSaperSternTwo} appeared, Saper and Stern noted that the result
7267: generalized to Satake compactifications of equal-rank symmetric spaces
7268: where all real boundary components were equal-rank---a proof replaced
7269: Lemmas ~11.6 and 11.7 from \cite{refnSaperSternTwo} with a case by case
7270: analysis based on the list appearing in \cite[(A2)]{refnZuckerLtwoIHTwo}.
7271: The proof here via Corollary~\ref{ssectBasicLemmaCor} and Lemma~
7272: \ref{ssectEqualRankLemma} is independent of classification theory.
7273: \end{rem*}
7274: 
7275: \section{Restriction to Fibers of $\pi\colon \Xhat \to \Xstar_\sigma$}
7276: \label{sectRestrictMicroSupportToFiber}
7277: 
7278: \begin{thm}
7279: \label{ssectRestrictMicroSupportToFiber}
7280: Let $F_R$ be a stratum of a Satake compactification $\Xstar_\sigma$ and
7281: assume that all the real boundary components $D_{R',h}\ge D_{R,h}$ of the
7282: associated Satake compactification $\lsb\RR\Dstar_\sigma$ are equal-rank.
7283: Let $x\in F_R$ and consider the inclusions $\ihat_{R,\l}\colon 
7284: \pi^{-1}(x)=\Xhat_{R,\l}\hookrightarrow \Xhat$ and $\ihat_R\colon \pi^{-1}(F_R) =
7285: X_R(L_{R,\l}) \hookrightarrow \Xhat$.  Let $\M$ be an $\L_{\Xhat}$-module.
7286: Then
7287: \begin{equation}
7288: \mS(\ihat_{R,\l}^*\M) \subseteq \Res_\l^* \ihat_R^*\mS(\M)
7289: \qquad\text{and}\qquad
7290: \mS(\ihat_{R,\l}^!\M) \subseteq \Res_\l^! \ihat_R^!\mS(\M)\ .
7291: \label{eqnMicrosupportRestrictionToPiInversePoint}
7292: \end{equation}
7293: 
7294: We have the estimates
7295: \begin{equation}
7296: \begin{split}
7297: d(\ihat_{R,\l}^*\M) \le \smash[b]{\sup_{\substack{\tilde V \in \mS(\M)\\ 
7298:  F_{\tilde R}\ge F_R}}}\bigl(&\, d(\tilde V;\M) + \frac12(\dim
7299:  D_{\tilde P,\l} + \dim D_{\tilde P,\l}(\tilde V)) \\
7300:  &\qquad + \frac12\codim_{F_{\tilde R}^*} F_R - \dim \sa_{\tilde P\cap
7301:  R}^{\tilde P} \,\bigr)
7302: \end{split}
7303: \label{eqnDREllStarEstimate}
7304: \end{equation}
7305: and 
7306: \begin{equation}
7307: \begin{split}
7308: c(\ihat_{R,\l}^!\M) \ge \smash[b]{\inf_{\substack{\tilde V\in \mS(\M)\\ F_{
7309:  \tilde R}\ge F_R}}} \bigl(&\, c(\tilde V;\M) - \frac12(\dim D_{\tilde P,\l} -
7310:  \dim D_{\tilde P,\l}(\tilde V)) \\
7311:  &\qquad + \frac12\codim_{F_{\tilde R}^*} F_R +
7312:  \dim \sa_{\tilde P\cap R}^{\tilde P} \bigr).
7313: \end{split}
7314: \label{eqnDREllShriekEstimate}
7315: \end{equation}
7316: In these estimates, $\tilde V$ is an irreducible $L_{\tilde P}$-module and
7317: we set $\tilde R = \tilde P^\dag$.
7318: \end{thm}
7319: 
7320: \begin{proof}
7321: Equation \eqref{eqnMicrosupportRestrictionToPiInversePoint} is simply
7322: Propositions ~\ref{ssectMicroSupportInverseImages} and
7323: \ref{ssectMicroSupportInclusionFiber}.  These propositions further imply
7324: that if $V_\l\in \mS(\ihat_{R,\l}^*\M)$, then there exists $V\in
7325: \mS(\ihat_R^*\M)$ and $\tilde V\in\mS(\M)$ such that $\Res_\l^*V=V_\l$ and
7326: there exists
7327: \begin{equation*}
7328: V=V_0\preccurlyeq_+
7329: \dots \preccurlyeq_+ V_N=\tilde V
7330: \end{equation*}
7331: satisfying \eqref{eqnRemoveHermitianRoot}--\eqref{eqnDominantCone}.
7332: Furthermore, as in the proof of Lemma ~\ref{ssectEqualRankTransferLemma},
7333: we have an almost direct product factorization
7334: \begin{equation*}
7335: L_{P_i}=\widetilde{L_{P_i,h}}L_{P_i,\l} = \widetilde{L_{P_i,h}}
7336: L_{P_i,\l}^{P_{i+1},h}L_{P_{i+1},\l} \ .
7337: \end{equation*}
7338: Let $D_{P_{i},\l}^{P_{i+1},h}$ denote the symmetric space associated to
7339: $L_{P_i,\l}^{P_{i+1},h}$.  Thus we can apply Corollary
7340: \ref{ssectEqualRankBasicLemmaCor} to the parabolic $(P_i/N_{P_{i+1}})\cap
7341: L_{P_{i+1},h} \subseteq L_{P_{i+1},h}$ and the induced Satake
7342: compactification of $D_{P_{i+1},h}$ in order to conclude that
7343: \begin{equation}
7344: [V_{i+1}:V_i] \le \frac12 ( \dim \n_{P_i}^{P_{i+1}} - \dim
7345: \sa_{P_i}^{P_{i+1}} - \dim D_{P_{i},\l}^{P_{i+1},h}(V_i) )
7346: \label{eqnOneStepDegreeBound}
7347: \end{equation}
7348: Now $\Res_{L_{P_{i+1},\l}}^{L_{P_i}}(V_i)$ is a direct sum of copies of
7349: $V_{P_{i+1},\l}$, so $D_{P_i,\l}(V_i) = D_{P_{i},\l}^{P_{i+1},h}(V_i)
7350: \times D_{P_{i+1},\l}(V_{i+1})$.  Consequently if we sum
7351: \eqref{eqnOneStepDegreeBound} for $i=0,\dots,N-1$ we obtain
7352: \begin{equation}
7353: \begin{split}
7354: [\tilde V: V] &\le \frac12 ( \dim \n_{P}^{\tilde P} - \dim \sa_{P}^{\tilde P}
7355: - \dim D_{P,\l}^{\tilde P,h}(V) )\\
7356: &=   - \frac12(\dim D_{P,\l}^{\tilde P,h} + \dim D_{P,\l}^{\tilde P,h}(V))
7357:      + \frac12\codim_{F_{\tilde R}^*} F_R 
7358:      - \dim \sa_{P}^{\tilde P}
7359: \end{split}
7360: \label{eqnReprDegreeEstimate}
7361: \end{equation}
7362: since
7363: \begin{equation*}
7364: \codim_{F_{\tilde R}^*} F_R = \dim \n_P^{\tilde P} + \dim \sa_P^{\tilde P}
7365: + \dim D_{P,\l}^{\tilde P,h}.
7366: \end{equation*}
7367: We now calculate that \eqref{eqnDREllStarEstimate} holds by using
7368: \eqref{eqndMDefn}, \eqref{eqnWeakMSInverseImageDegree},
7369: \eqref{eqnDegreeFiberRestrictionEstimates}, and
7370: \eqref{eqnReprDegreeEstimate}.  The proof of \eqref{eqnDREllShriekEstimate}
7371: is similar.
7372: \end{proof}
7373: 
7374: \begin{cor}
7375: \label{ssectRestrictMicroSupportToFiberCorollary}
7376: In the above situation, assume further that $\emS(\M)=\{E\}$ for
7377: $E\in\IrrRep(G)$ with $c(E;\M)=d(E;\M)=0$.  Then
7378: \begin{equation*}
7379: d(\ihat_{R,\l}^*\M) \le \frac12\codim F_R - \dim \sa_{R}^{G}
7380: \qquad\text{and}\qquad c(\ihat_{R,\l}^!\M) \ge \frac12\codim F_R + \dim
7381: \sa_{R}^{G}.
7382: \end{equation*}
7383: \end{cor}
7384: 
7385: \begin{proof}
7386: Any $V\in \mS(\M)$ satisfies $V\preccurlyeq_0 E$ by Corollary
7387: ~\ref{ssectPartialOrderingOnMicroSupport} and therefore has the form
7388: $H^{\l(w)}(\n_{P};E)_w$ where $w$ is fundamental by Lemma
7389: ~\ref{ssectPartialOrdering}.  However since $D$ is equal-rank the only
7390: fundamental parabolic $\RR$-subgroup is $G$.  Hence $\mS(\M)=\{E\}$.  Now
7391: apply \eqref{eqnDREllStarEstimate} and \eqref{eqnDREllShriekEstimate}.
7392: \end{proof}
7393: 
7394: \section{Application: the Conjecture of Rapoport and Goresky-MacPherson}
7395: \label{sectRapoportConjecture}
7396: 
7397: \begin{thm}
7398: \label{ssectRapoportConjecture}
7399: Let $\Xstar$ be a Satake compactification of $X$ for which all real
7400: boundary components $D_{R,h}$ are equal-rank, and let $\pi\colon \Xhat \to
7401: \Xstar$ be the projection from the reductive Borel-Serre compactification.
7402: Let $p$ be a middle perversity.  Then there is a natural quasi-isomorphism
7403: $\pi_*\IpC(\Xhat;\EE) \cong \IpC(\Xstar;\EE)$.
7404: \end{thm}
7405: \begin{rem*}
7406: When $D$ is Hermitian and $\Xstar$ is the Baily-Borel-Satake
7407: compactification, the equal-rank hypothesis is automatic.  In this case the
7408: theorem was conjectured independently by Rapoport
7409: \cite{refnRapoportLetterBorel} and Goresky and MacPherson
7410: \cite{refnGoreskyMacPhersonWeighted}; a proof in this case for $\QQrank
7411: G=1$ was given by Saper and Stern \cite[Appendix]{refnRapoport}.
7412: \end{rem*}
7413: 
7414: \begin{proof}
7415: For $x\in F_R$, a proper stratum of $\Xstar$, let $\i_x\colon
7416: \{x\}\hookrightarrow \Xstar$ and $\ihat_{R,\l}\colon \pi^{-1}(x) =
7417: \Xhat_{R,\l}\hookrightarrow \Xhat$ be the inclusion maps.  Let $q(k)= k -2
7418: - p(k)$ be the dual middle perversity.  By the local characterization of
7419: intersection cohomology \cite{refnGoreskyMacPhersonIHTwo}, \cite[V,
7420: 4.2]{refnBorelIntersectionCohomology} on $\Xstar$ we need to verify that
7421: \begin{enumerate}
7422: \item $\pi_*\IpC(\Xhat;\EE)$ is $\X$-clc,
7423: \label{itemIHCLC}
7424: \item $\pi_*\IpC(\Xhat;\EE)|_X$ is quasi-isomorphic to $\EE$,
7425: \label{itemIHOpenStratum}
7426: \item $H^i(\i_x^*\pi_*\IpC(\Xhat;\EE)) = 0$ for $x\in F_R$, $i> p(\codim
7427: F_R)$, and
7428: \label{itemIHVanishing}
7429: \item $H^i(\i_x^!\pi_*\IpC(\Xhat;\EE)) = 0$ for $x\in F_R$, $i< \dim X -
7430: q(\codim F_R)$.
7431: \label{itemIHCovanishing}
7432: \end{enumerate}
7433: Condition ~\itemref{itemIHCLC} follows from \cite[V,
7434: 10.16]{refnBorelIntersectionCohomology} and condition
7435: ~\itemref{itemIHOpenStratum} is obvious.  For the others, calculate
7436: \begin{alignat*}{2}
7437: H(\i_x^*\pi_*\IpC(\Xhat;\EE)) &\cong
7438: H(\Xhat_{R,\l};\ihat_{R,\l}^*\IpC(\Xhat;\EE)) &\qquad
7439: &\text{by \cite[V, 10.7]{refnBorelIntersectionCohomology},} \\
7440: &\cong H(\Xhat_{R,\l};\ihat_{R,\l}^*\IpC(E)) &\qquad&\text{by Thm.
7441: ~\ref{ssectRealizationLModules} and Prop.
7442: ~\ref{ssectIntersectionCohomology}}
7443: \end{alignat*}
7444: and similarly $H(\i_x^*\pi_*\IpC(\Xhat;\EE)) \cong
7445: H(\Xhat_{R,\l};\ihat_{R,\l}^!\IpC(E))$.  Now since all strata of $\Xstar$
7446: have even dimension by the equal-rank hypothesis, $p(\codim F_R) =
7447: \frac12\codim F_R - 1$ and $\dim X - q(\codim F_R)= \frac12\codim F_R + 1$.
7448: Thus we need to demonstrate that
7449: \begin{equation}
7450: \begin{alignedat}{2}
7451: H^i(\Xhat_{R,\l};\ihat_{R,\l}^*\IpC(E))&=0  &\qquad&\text{for $i\ge
7452:   \frac12\codim F_R$,}  \\ 
7453: H^i(\Xhat_{R,\l};\ihat_{R,\l}^!\IpC(E))&=0  &\qquad&\text{for $i\le
7454:   \frac12\codim F_R$.}
7455: \end{alignedat}
7456: \label{eqnVanishingCoVanishing}
7457: \end{equation}
7458: 
7459: Under the equal-rank hypotheses, the possible types of irreducible
7460: $\QQ$-root systems that can occur in $G$ are $B_n$, $BC_n$, $C_n$, and
7461: $G_2$ \cite[(A.2)]{refnZuckerLtwoIHTwo}.  Thus we can apply Corollary
7462: ~\ref{ssectIHMicroPurityCorollary} to see that the essential micro-support
7463: of $\IpC(E)$ is simply $\{E\}$; Corollary
7464: ~\ref{ssectRestrictMicroSupportToFiberCorollary} then implies that
7465: $d(\ihat_{R,\l}^*\M) < \frac12\codim F_R$ and $c(\ihat_{R,\l}^!\M) >
7466: \frac12\codim F_R$.  The proof \eqref{eqnVanishingCoVanishing} is concluded
7467: by applying our vanishing theorem for the cohomology of $\L$-modules,
7468: Theorem ~\ref{ssectGlobalVanishing}.
7469: \end{proof}
7470: 
7471: \subsection{}
7472: \label{ssectGoreskyHarderMacPhersonTheorem}
7473: If we replace the use of Corollary ~\ref{ssectIHMicroPurityCorollary} by
7474: Theorem ~\ref{ssectWeightCohomologyMicroSupport} in the preceding proof we
7475: obtain the following theorem.
7476: 
7477: \begin{thm*}
7478: Let $\Xstar$ be a Satake compactification of $X$ for which all real
7479: boundary components $D_{R,h}$ are equal-rank, and let $\pi\colon \Xhat \to
7480: \Xstar$ be the projection from the reductive Borel-Serre compactification.
7481: Let $\eta$ be a middle weight profile and let $p$ be a middle perversity.
7482: Then there is a natural quasi-isomorphism $\pi_*\WnC(\Xhat;\EE) \cong
7483: \IpC(\Xstar;\EE)$.
7484: \end{thm*}
7485: \begin{rem*}
7486: When $D$ is Hermitian and $\Xstar$ is the Baily-Borel-Satake
7487: compactification, this theorem was the main result of
7488: \cite{refnGoreskyHarderMacPherson}.
7489: \end{rem*}
7490: 
7491: %%% References
7492: 
7493: \bibliographystyle{amsplain}
7494: \bibliography{rbs}
7495: \end{document}
7496: