1: \documentclass[11pt,final]{ucthesis}
2: \usepackage{amsmath}
3: \usepackage{amssymb}
4: \usepackage{amsfonts}
5: \usepackage{fancyheadings}
6: \usepackage{amscd}
7: \usepackage{latexsym}
8: \usepackage{theorem}
9: % \usepackage{doublespace}
10: \usepackage{epsfig}
11: \usepackage{psfrag}
12: \usepackage{bbold}
13:
14: % \include{ltxtemp_th}
15:
16: \newtheorem{theorem}{Theorem}[chapter]
17: \newtheorem{lemma}[theorem]{Lemma}
18: \newtheorem{proposition}[theorem]{Proposition}
19: \newtheorem{corollary}[theorem]{Corollary}
20: \newtheorem{claim}[theorem]{Claim}
21: \newtheorem{conjecture}[theorem]{Conjecture}
22:
23: {\theorembodyfont{\rmfamily}
24: \theoremstyle{plain}
25: \newtheorem{definition}[theorem]{Definition}
26: \newtheorem{convention}[theorem]{Notational Convention}
27: \newtheorem{example}[theorem]{Example}
28: \newtheorem{problem}[theorem]{Problem}
29: \newtheorem{remark}[theorem]{Remark}
30: }
31:
32: \input{xy}
33: \xyoption{all}
34:
35: % Margin stuff from jason
36: % \oddsidemargin=0pt
37: % \evensidemargin=0pt
38: % \topmargin=0in
39: \headheight=14pt
40: % \headsep=0pt
41: % \setlength{\textheight}{9in}
42: % \setlength{\textwidth}{6.5in}
43:
44: \newcommand{\Vol}{\operatorname{Vol}}
45: \renewcommand{\Re}{\operatorname{Re}}
46: \renewcommand{\Im}{\operatorname{Im}}
47: \newcommand{\im}{\operatorname{im}}
48: \newcommand{\id}{\operatorname{id}}
49: \newcommand{\Coker}{\operatorname{Coker}}
50: \renewcommand{\ker}{\operatorname{Ker}}
51: \renewcommand{\dim}{\operatorname{dim}}
52: \newcommand{\Ann}{\operatorname{Ann}}
53: \newcommand{\Sym}{\operatorname{Sym}}
54: \newcommand{\Hom}{\operatorname{Hom}}
55: \newcommand{\tr}{\operatorname{tr}}
56:
57: \newcommand{\Ld}{\displaystyle\lim}
58: \newcommand{\Od}{\displaystyle\bigoplus}
59: \newcommand{\Pd}{\displaystyle\prod}
60: \newcommand{\Sd}{\displaystyle\sum}
61:
62: \newcommand{\C}{{\mathbb{C}}}
63: \newcommand{\Z}{{\mathbb{Z}}}
64: \newcommand{\Q}{{\mathbb{Q}}}
65: \renewcommand{\H}{{\mathbb{H}}}
66: \newcommand{\R}{{\mathbb{R}}}
67: \newcommand{\Zt}{{\Z_2}}
68:
69: \newcommand{\otn}{\{1,\ldots,n\}}
70: \newcommand{\half}{\frac{1}{2}}
71: \newcommand{\bd}{\partial}
72:
73: \newcommand{\bigmid}{\hs\Big{|}\hs}
74: \newcommand{\subs}{\subseteq}
75: \newcommand{\sups}{\supseteq}
76: \newcommand{\hookto}{{\hookrightarrow}}
77: \newcommand{\onto}{{\twoheadrightarrow}}
78: \renewcommand{\iff}{\Leftrightarrow}
79: \newcommand{\impl}{\Rightarrow}
80: \newcommand{\becircled}{\mathaccent "7017}
81: \newcommand{\hs}{\hspace{3pt}}
82: \renewcommand{\ss}{\substack}
83:
84: \newcommand{\D}{\Delta}
85: \renewcommand{\i}{\iota}
86: \renewcommand{\a}{\alpha}
87: \renewcommand{\b}{\beta}
88: \newcommand{\s}{\sigma}
89: \newcommand{\eps}{\varepsilon}
90: \renewcommand{\k}{\kappa}
91: \newcommand{\K}{\kappa}
92: \newcommand{\ga}{\gamma}
93: \renewcommand{\t}{\tau}
94:
95: \renewcommand{\cot}{T^*\C^n}
96: \newcommand{\Cn}{\C^n}
97: \newcommand{\Cno}{\C^{n+1}}
98:
99: \newcommand{\Mso}{\Ma^{\so}}
100: \newcommand{\Macs}{\Ma^{\gm}}
101: \newcommand{\Ma}{\M}
102: \newcommand{\Xa}{\X}
103: \newcommand{\Gap}{\Gamma_p}
104: \newcommand{\Gapt}{\widehat{\Gamma}_p}
105: \newcommand{\Gapto}{\Gapt\!\becircled{}}
106:
107: \newcommand{\muc}{\mu_{\C}}
108: \newcommand{\mur}{\mu_{\R}}
109: \newcommand{\barmuc}{\bar{\mu}_{\C}}
110: \newcommand{\barmur}{\bar{\mu}_{\R}}
111: \newcommand{\mr}{\mur}
112: \newcommand{\mc}{\muc}
113: \newcommand{\mhk}{\mu_{\text{HK}}}
114: \newcommand{\mh}{\mr\oplus\mc}
115: \newcommand{\ka}{K_{\a}}
116: \newcommand{\hka}{\hk_{\a}}
117: \newcommand{\hk}{H\!K}
118: \newcommand{\htk}{H^*_{T^k}(pt)}
119: \renewcommand{\mod}{{\!/\!\!/}}
120: \newcommand{\mmod}{{\!/\!\!/\!\!/\!\!/}}
121:
122: \newcommand{\SO}{{SO(3)}}
123: \newcommand{\SL}{{SL(2,\C)}}
124: \newcommand{\sutd}{{\mathfrak{su}(2)}^*}
125: \newcommand{\sut}{{\mathfrak{su}(2)}}
126: \newcommand{\slt}{{\mathfrak{sl}(2,\C)}}
127: \newcommand{\so}{S^1}
128: \newcommand{\gm}{\C^{\!\times}}
129:
130: \newcommand{\GC}{G_{\C}}
131: \newcommand{\gdc}{\gd_{\C}}
132: \newcommand{\gd}{\g^*}
133: \newcommand{\g}{\mathfrak{g}}
134: \newcommand{\Tk}{T^k}
135: \newcommand{\Tn}{T^n}
136: \newcommand{\Td}{T^d}
137: \newcommand{\Tkd}{(\tk_{\Z})^*}
138: \newcommand{\Tnd}{(\tn_{\Z})^*}
139: \newcommand{\Tdd}{(\td_{\Z})^*}
140: \newcommand{\tk}{\mathfrak{t}^k}
141: \newcommand{\tn}{\mathfrak{t}^n}
142: \newcommand{\td}{\mathfrak{t}^d}
143: \newcommand{\tkd}{(\tk)^*}
144: \newcommand{\tndz}{(\tn_{\Z})^*}
145: \newcommand{\tnz}{\tn_{\Z}}
146: \newcommand{\tdz}{\td_{\Z}}
147: \newcommand{\tnd}{(\tn)^*}
148: \newcommand{\tdd}{(\td)^*}
149: \newcommand{\tdu}{\mathfrak{t}^*}
150:
151: \newcommand{\E}{\mathcal{E}}
152: \newcommand{\EA}{\E_A}
153: \renewcommand{\L}{\mathcal{L}}
154:
155: \newcommand{\Af}{\operatorname{Rep}(Q)}
156: \newcommand{\Diag}{\operatorname{Diag}}
157:
158: % ht commands
159: \newcommand{\ktd}{\k_{\Td}}
160: \newcommand{\ktdso}{\k_{\Td\times\so}}
161: \newcommand{\kso}{\k_{\so}}
162: \newcommand{\otd}{\{1,\ldots,d\}}
163: \newcommand{\Tdrs}{T_{\R}^d\times\Zt}
164: \newcommand{\Tdr}{T_{\R}^d}
165: \newcommand{\Tdso}{\Td\times S^1}
166: \newcommand{\Tnso}{\Tn\times S^1}
167: \newcommand{\hma}{H^*(\Ma)}
168: \newcommand{\htm}{H^*_{\Td}(\Ma)}
169: \newcommand{\htsm}{H^*_{\Td\times S^1}(\Ma)}
170: \newcommand{\hsm}{H^*_{S^1}(\Ma)}
171: \newcommand{\hr}{H^*_{\Tdr}(\M_{\R};\Zt)}
172: \newcommand{\hrs}{H^*_{\Tdr\times\Zt}(\M_{\R};\Zt)}
173: \newcommand{\hscomp}{H^*_{\Zt}(\MA;\Zt)}
174: \newcommand{\hcomp}{H^*(\MA;\Zt)}
175: \newcommand{\bomp}{\C^d\setminus\cup_{i=1}^n H_i^{\C}}
176: \newcommand{\os}{\mathcal{OS}}
177: \newcommand{\ost}{\mathcal{OS}\otimes\Zt}
178:
179: % os commands
180: \newcommand{\A}{\mathcal{A}}
181: \newcommand{\MA}{\mathcal{M}(\A)}
182: \newcommand{\hz}{H_{\Zt}^*(pt)}
183:
184: % hp commands
185: \newcommand{\ctn}{{\C^{2n}}}
186: \newcommand{\ses}{st}
187: \newcommand{\Ae}{[A;e_1,\ldots,e_n]}
188: \newcommand{\pathsp}{\prod_{i=1}^n S^2_{\a_i}}
189: \newcommand{\polsa}{\X_S}
190: \newcommand{\pola}{\X}
191: \newcommand{\mcs}{\mathcal{S}}
192: \newcommand{\sa}{\mcs}
193: \newcommand{\spa}{\mcs'}
194: \newcommand{\core}{\mathfrak{L}}
195: \newcommand{\corea}{\mathfrak{L}_{\a}}
196: \newcommand{\zns}{z_{\ns}
197: \newcommand{\zms}{z_{\ms}}}
198: \newcommand{\ms}{m_S}
199: \newcommand{\ns}{n_S}
200: \newcommand{\bi}{d_i}
201: \newcommand{\bk}{d_k}
202: \newcommand{\bj}{d_j}
203: \newcommand{\ba}{d_A}
204: \newcommand{\bns}{d_{\ns}}
205: \newcommand{\ci}{c_i}
206: \newcommand{\ck}{c_k}
207: \newcommand{\cns}{c_{\ns}}
208: \newcommand{\Sb}{\bar{S}}
209: \newcommand{\Lb}{\overline{S^c}}
210: \newcommand{\Ac}{A^c}
211: \newcommand{\asb}{A\cap\Sb}
212: \newcommand{\acsb}{\Ac\cap\Sb}
213: \newcommand{\pis}{\prod_{i\in\Sb}}
214: \newcommand{\pjl}{\prod_{j\in\Lb}}
215: \newcommand{\vs}{v_S}
216: \newcommand{\vst}{\tilde{v}_S}
217: \newcommand{\twn}{\{2,\ldots,n\}}
218: \newcommand{\Zb}{\Z[d_1,\ldots,d_n]/\langle\bk^2-d_1\bk\mid k=2,\ldots,n\rangle}
219: \newcommand{\Qb}{\Q[d_1,\ldots,d_n]/\langle\bk^2-d_1\bk\mid k=2,\ldots,n\rangle}
220: \newcommand{\us}{U_S}
221: \newcommand{\mtus}{X_T\cap\us}
222: \newcommand{\ts}{T\sups S}
223: \newcommand{\sz}{|S|-1}
224:
225: % hm commands
226: \newcommand{\Kh}{\hat\k}
227: \newcommand{\Hn}{\H^n}
228: \newcommand{\rgt}{r_T^G}
229: \newcommand{\phiw}{\phi^W}
230: \newcommand{\phat}{\hat{\phi}}
231: \newcommand{\phatw}{\hat{\phi}^W}
232: \newcommand{\IM}{\int_M}
233: \newcommand{\IN}{\int_N}
234: \newcommand{\IG}{\int_G}
235: \newcommand{\IF}{\int_F}
236: \newcommand{\intxt}{\int_{X\mmod T}}
237: \newcommand{\intxg}{\int_{X\mmod G}}
238: \newcommand{\I}{\mathcal{I}}
239: \newcommand{\J}{\mathcal{J}}
240: \newcommand{\imkt}{\left(\Im\K_T\right)}
241: \newcommand{\hso}{H_{\so}^*}
242: \newcommand{\hhso}{\widehat{H}_{\so}^*}
243: \newcommand{\hmg}{H_{\so}^*(M\mmod G)}
244: \newcommand{\hmt}{H_{\so}^*(M\mmod T)}
245: \newcommand{\hxg}{H_{\so}^*(X\mmod G)}
246: \newcommand{\hxt}{H_{\so}^*(X\mmod T)}
247: \newcommand{\hhxg}{\widehat{H}_{\so}^*(X\mmod G)}
248: \newcommand{\hhxt}{\widehat{H}_{\so}^*(X\mmod T)}
249: \newcommand{\hhst}{\widehat{H}_{\so\times T}^*(X)}
250: \newcommand{\hhsg}{\widehat{H}_{\so\times G}^*(X)}
251: \newcommand{\hp}{H_{S^1}^*(pt)}
252: \newcommand{\hhp}{{\mathbb K}}
253: \newcommand{\hm}{\hso(M)}
254: \newcommand{\hhn}{\hhso(N)}
255: \newcommand{\hhm}{\hhso(M)}
256: \newcommand{\X}{{\mathfrak{X}}}
257: \newcommand{\M}{{\mathfrak{M}}}
258: % \newcommand{\M}{M}
259: \newcommand{\MM}{{\M\times\M}}
260: \newcommand{\NN}{\mathfrak{N}}
261: \newcommand{\dso}{\Diag_*(1)}
262: \newcommand{\muh}{\mu_{G}}
263: \newcommand{\mubh}{\mu_{T}}
264: \newcommand{\bigmod}{{\Big/\!\!\!\!\Big/}}
265: \newcommand{\bigmmod}{{\Big/\!\!\!\!\Big/\!\!\!\!\Big/\!\!\!\!\Big/}}
266:
267: % kernel commands
268: \newcommand{\ra}{\rangle}
269: \newcommand{\Hir}{H^r_i}
270: \newcommand{\PrA}{P^r_{\! A}}
271: \renewcommand{\Pr}{P^r}
272: \newcommand{\Dr}{\D^r}
273: \newcommand{\DrA}{\D^r_A}
274: \newcommand{\DrzA}{\D^{r_0}_A}
275: \newcommand{\rt}{\tilde r}
276: \newcommand{\odra}{\mathbf{1}_{\DrA}}
277: \newcommand{\odr}{\mathbf{1}_{\Dr}}
278: \newcommand{\WA}{W_A}
279: \newcommand{\WAbd}{W_A^{bd}}
280: \newcommand{\Fbd}{\mathcal{F}^{bd}}
281: \newcommand{\F}{\mathcal{F}}
282:
283: \newcommand{\<}{\left<}
284: \renewcommand{\>}{\right>}
285: \newcommand{\la}{\langle}
286: \renewcommand{\ra}{\rangle}
287: \renewcommand{\(}{\left(}
288: \renewcommand{\)}{\right)}
289:
290: \newcommand{\qed}{\hfill \mbox{$\Box$}\medskip\newline}
291: \newenvironment{proof}{\noindent {\bf Proof:}}{\qed \par}
292: \newenvironment{noproof}{\noindent {\bf Proof:}}{\newline}
293: \newenvironment{sketch}{\noindent {\bf Sketch of Proof:}}{\qed \par}
294:
295: \newenvironment{proofcomb}{\noindent {\bf Proof of \ref{comb}:}}{\qed\par}
296: \newenvironment{proofint}{\noindent {\bf Proof of \ref{int}:}}{}
297: \newenvironment{proofintegration}{\noindent {\bf Proof of \ref{integration}:}}{\qed \par}
298: \newenvironment{proofordinary}{\noindent {\bf Proof of \ref{ordinary}:}}{\qed \par}
299: \newenvironment{proofhtsm}{\noindent {\bf Proof of \ref{htsm}:}}{\qed \par}
300: \newenvironment{proofp}{\noindent {\bf Proof of \ref{placement}:}}{\qed \par}
301: \newenvironment{proofc}{\noindent {\bf Proof of \ref{coorientation}:}}{\qed \par}
302: \newenvironment{proofhsm}{\noindent {\bf Proof of \ref{hsm}:}}{\qed \par}
303: \newenvironment{proofmain}{\noindent {\bf Proof of \ref{main}:}}{\qed \par}
304: \newenvironment{prooftech}{\noindent {\bf Proof of \ref{tech}:}}{\qed \par}
305: \newenvironment{componentproof}{\noindent {\bf Proof of \ref{component}:}}{\qed \par}
306: \newenvironment{usproof}{\noindent {\bf Proof of \ref{us}:}}{\qed \par}
307: \newenvironment{mainproof}{\noindent {\bf Proof of \ref{main}:}}{\qed \par}
308: \newenvironment{eqcoreproof}{\noindent {\bf Proof of \ref{eqcore}:}}{\qed \par}
309:
310: % \newcommand{\thisisstupid}{\S}
311: % \makeatletter
312: % \renewcommand{\@seccntformat}[1]{$\thisisstupid$
313: % \csname the#1\endcsname\hspace{0.5em}}
314: % \makeatother
315:
316: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
317: %
318: % Double spacing, if you want it.
319: %
320: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
321: % \def\dsp{\def\baselinestretch{2.0}\large\normalsize}
322: % \dsp
323:
324: % \addtolength{\voffset}{+0.1in}
325: % %\addtolength{\hoffset}{+0.5in}
326: % \addtolength{\textwidth}{-0.2in}
327: % \addtolength{\textheight}{-0.4in}
328:
329: \hyphenation{mar-gin-al-ia}
330:
331: \begin{document}
332:
333: \title{Hyperk\"ahler Analogues of K\"ahler Quotients}
334: \author{Nicholas James Proudfoot}
335: \degreesemester{Spring}
336: \degreeyear{2004}
337: \degree{Doctor of Philosophy}
338: \chair{Professor Allen Knutson}
339: \othermembers{Professor Robion Kirby \\
340: Professor Robert Littlejohn}
341: \numberofmembers{3}
342: \prevdegrees{A.B. (Harvard University) 2000}
343: \field{Mathematics}
344: \campus{Berkeley}
345:
346: \maketitle
347: \approvalpage
348: \copyrightpage
349: \pagestyle{fancyplain}
350: % \include{abstract}
351:
352: \begin{abstract}
353: Let $\X$ be a K\"ahler manifold that is presented as a K\"ahler
354: quotient of $\Cn$ by the linear action of a compact group $G$.
355: We define the {\em hyperk\"ahler analogue} $\M$ of $\X$ as a
356: hyperk\"ahler quotient of the cotangent bundle $\cot$ by the induced
357: $G$-action. Special instances of this construction include
358: hypertoric varieties \cite{BD,K1,HS,HP1} and quiver varieties \cite{N1,N2,N3}.
359: One of our aims is to provide a unified treatment of these
360: two previously studied examples.
361:
362: The hyperk\"ahler analogue $\M$ is noncompact, but this noncompactness
363: is often ``controlled'' by an action of $\gm$ descending from
364: the scalar action on the fibers of $\cot$. Specifically, we are
365: interested in the case where the moment map for the action
366: of the circle $\so\subs\gm$
367: is proper. In such cases, we
368: define the {\em core} of $\M$, a reducible, compact subvariety onto which
369: $\M$ admits a circle-equivariant deformation retraction. One of the components
370: of the core is isomorphic to the original K\"ahler manifold $\X$.
371: % When $\X$ is a toric variety, every other core component is also a toric variety.
372: When $\X$ is a moduli space of polygons in $\R^3$, we interpret
373: each of the other core components of $\M$ as related polygonal moduli spaces.
374:
375: Using the circle action with proper moment map, we define an integration
376: theory on the circle-equivariant cohomology of $\M$,
377: motivated by the well-known localization theorem
378: of \cite{AB} and \cite{BV}. This allows us to prove a hyperk\"ahler
379: analogue of Martin's theorem \cite{Ma}, which describes the cohomology
380: ring of an arbitrary K\"ahler quotient in terms of the
381: cohomology of the quotient by a maximal torus.
382: This theorem, along with a direct analysis of the equivariant cohomology
383: ring of a hypertoric variety, gives us a method for computing
384: the equivariant cohomology ring of many hyperk\"ahler analogues,
385: including all quiver varieties.
386: \abstractsignature
387: \end{abstract}
388:
389: \pagestyle{plain}
390: \begin{frontmatter}
391: \pagenumbering{roman}
392: % \begin{dedication}
393: % \null\vfil
394: % {\large
395: % \begin{center}
396: % To the fond memory of a great teacher, \\
397: % Edwin L. Shay
398: % \end{center}}
399: % \vfil\null
400: % \end{dedication}
401:
402: \tableofcontents
403: %\addcontentsline{toc}{sec_name}{Chapter 1}
404: %\listoffigures
405: %\listoftables
406: \lhead[\fancyplain{}{}]{\fancyplain{}{}}
407: \rhead[\fancyplain{}{}]{\fancyplain{}{}}
408:
409: \begin{acknowledgements}
410: During my years at Berkeley, I have had countless invaluable conversations
411: with other students, post-docs, professors, and visitors to Berkeley and MSRI.
412: Many of these conversations
413: have been relevant to the work described here, and many more
414: have greatly influenced my general mathematical perspective.
415: I would like to particularly acknowledge my coauthors, Megumi Harada
416: and Tam\'as Hausel, and my advisor, Allen Knutson.
417: I feel extremely lucky to be part of such a warm and lively community.
418: \end{acknowledgements}
419:
420: \end{frontmatter}
421:
422: \pagestyle{fancy}
423: \pagenumbering{arabic}
424: \setlength{\headrulewidth}{1pt}
425: \lhead[\rm\thepage]{\sl\leftmark}
426: \rhead[\sl\leftmark]{\rm\thepage}
427: \renewcommand{\chaptermark}[1]{\markboth{\chaptername\hs\thechapter. #1}{ }}
428: \cfoot{}
429: % \renewcommand{\chaptermark}[1]{\markboth{#1}{ }}
430: % \rhead[\fancyplain{}{\bfseries \leftmark}]{\fancyplain{}{\bfseries \thepage}}
431:
432: \begin{chapter}{Introduction}
433: % A {\em K\"ahler} manifold is a riemannian manifold $(X,g)$ with a parallel,
434: % orthogonal complex structure $J$. These conditions imply that the $2$-form
435: % $\omega(v,w) := g(Jv,w)$ is closed and nondegenerate,
436: % hence a K\"ahler manifold may be thought of
437: % as a manifold which is simultaneously riemannian,
438: % complex, and symplectic, in a compatible way.
439: % Interesting examples of K\"ahler manifolds include
440: % toric varieties \cite{Do}, polygon spaces \cite{HK},
441: % and the moduli space of stable vector bundles on a curve \cite{NS}.
442: We begin with a quick overview of some of the structures that we will consider
443: in this thesis, and the types of questions that we will be asking.
444: Detailed definitions will be deferred until the next chapter.
445:
446: Let $G$ be a compact Lie group acting linearly on $\Cn$,
447: with moment map $\mu:\Cn\to\gd$, and suppose we are given
448: a central (i.e. $G$-fixed) regular value $\a\in\gd$ of $\mu$.
449: From this data, we may define the K\"ahler
450: quotient $$\X := \Cn\mod G = \mu^{-1}(\a)/G,$$ which itself inherits the structure of a
451: K\"ahler manifold.
452: % (If the complexification $G^{\C}$
453: % of $G$ is algebraic, and acts algebraically
454: % on $\Cn$, then $\X$ may be thought of as
455: (We may also think of $\X$ as the geometric invariant theory
456: quotient of $\Cn$ by $G^{\C}$ in the sense of Mumford \cite[\S 8]{MFK};
457: see Proposition \ref{hkhs}.)
458: % Toric varieties arise from the case where $G$ is abelian, and polygon spaces
459: % arise when the vector space in question is the space of representations of a
460: % certain quiver (see Section \ref{quiver}).
461: % The moduli space of stable vector bundles on a curve arises from an analogous
462: % construction in which $\Cn$ is replaced by the infinite-dimensional affine space
463: % of holomorphic structures on a fixed smooth bundle $E$, which is acted on by
464: % the gauge group of automorphisms of $E$.
465: A {\em hyperk\"ahler} manifold is a riemannian manifold $(M,g)$ equipped
466: with three orthogonal complex structures $J_1, J_2, J_3$ and three compatible symplectic
467: forms $\omega_1, \omega_2, \omega_3$ such that $J_1J_2= -J_2J_1 = J_3$
468: for $i=1,2$, and $3$.
469: The cotangent bundle $\cot$ has a natural hyperk\"ahler structure,
470: and this structure is preserved by the induced action of $G$.
471: Furthermore, there exist maps $\mu_i:\cot\to\gd$ for $i=1,2,3$
472: such that $\mu_i$ is a moment map for the action of $G$ with respect
473: to the symplectic form $\omega_i$.
474: We define the {\em hyperk\"ahler analogue} of $\X$
475: to be the hyperk\"ahler quotient
476: $$\M := \cot\mmod G = \Big(\mu_1^{-1}(\a)\cap\mu_2^{-1}(0)
477: \cap\mu_3^{-1}(0)\Big)\Big/G.$$
478: (The set
479: $\mu_2^{-1}(0)\cap\mu_3^{-1}(0)\subs\cot$
480: is a complex subvariety with respect to $J_1$, and $\M$
481: may be thought of as the geometric invariant theory quotient
482: of this variety by $G^{\C}$.)
483: The quotient $\M$ is a complete hyperk\"ahler manifold \cite{HKLR},
484: containing $T^*\X$ as a dense open subset (see Proposition \ref{denseopen}).
485: The following is a description of some well-known classes of K\"ahler quotients,
486: along with their hyperk\"ahler analogues.
487:
488: \vspace{.2cm}
489: \noindent{\bf Toric and hypertoric varieties.}
490: These examples comprise the case where $G$ is abelian.
491: The geometry of toric varieties is deeply related to the combinatorics of polytopes;
492: for example, Stanley \cite{St} used the hard Lefschetz theorem for toric varieties
493: to prove certain inequalities for the $h$-numbers of a simplicial polytope.
494: Hypertoric varieties, introduced by Bielawski and Dancer \cite{BD}, interact in a similar
495: way with the combinatorics of real hyperplane arrangements.
496: Following Stanley's work, Hausel and Sturmfels \cite{HS} used the hard Lefschetz theorem
497: on a hypertoric variety to give a geometric interpretation of some previously-known
498: inequalities for the $h$-numbers of a rationally representable matroid.
499: In Chapter \ref{ht} we will explore further combinatorial properties
500: of the various equivariant cohomology rings of a hypertoric variety.
501:
502: \vspace{.2cm}
503: \noindent{\bf Quiver varieties.}
504: For any directed graph,
505: Nakajima \cite{N1,N2,N3} defined a {\em quiver variety}
506: to be the hyperk\"ahler analogue of the moduli space of framed representations of that graph
507: (see Section \ref{quiver}).
508: Examples include the Hilbert scheme of $n$ points in $\C^2$ \cite{N4}, the moduli space
509: of instantons on an ALE space \cite{N1}, and Konno's hyperpolygon spaces \cite{K2,HP2},
510: which are the hyperk\"ahler analogues of the moduli spaces of $n$-sided polygons in $\R^3$ with
511: fixed edge lengths.
512: Quiver varieties have received much attention from representation theorists due to the actions
513: of various infinite-dimensional Lie algebras on the cohomology and
514: K-theory of a quiver variety (see, for example, \cite{N3,N5}).
515:
516: \vspace{.2cm}
517: \noindent{\bf Moduli spaces of bundles and connections.}
518: Narasimhan and Seshadri \cite{NS} defined a notion of stability for a vector bundle
519: on a Riemann surface $\Sigma$, and proved that the moduli space of stable
520: holomorphic bundles on $\Sigma$ may be identified with the moduli space of irreducible,
521: flat, unitary connections.
522: Atiyah and Bott presented this space as a K\"ahler quotient of the affine
523: space of all connections on a fixed bundle $E$ by the gauge group
524: of automorphisms of $E$. This picture can be complexified by replacing holomorphic
525: bundles with Higgs bundles, and unitary connections with arbitrary ones \cite{Hi}.
526: The correspondence between Higgs bundles, flat connections, and representations
527: of the fundamental group is known as {\em nonabelian Hodge theory}, and has
528: been studied and generalized extensively by Simpson \cite{Si} in addition to
529: many other authors.
530: These constructions involve taking a quotient of an infinite dimensional affine
531: space by an infinite dimensional group, and therefore lie
532: it is beyond the scope of this work.
533: Many of our techniques, however, can be applied in this context.
534: See for example \cite{H1, HT1, HT2}.
535: % A Higgs bundle on a Riemann surface $\Sigma$ is a pair $(E,\theta)$, where
536: % $E$ is a holomorphic vector bundle on $\Sigma$, and $\Phi$ is an $\operatorname{End}(E)$
537: % valued $1$-form of fixed determinant. Hitchin \cite{Hi} generalized the notion
538: % of stability of vector bundles to Higgs bundles,
539: % and interpreted the moduli space of stable rank $2$ Higgs bundles on $\Sigma$
540: % as the moduli space of self-dual solutions to the Yang-Mills equations.
541: % In arbitrary rank, Simpson \cite{Si} showed that
542: % the moduli space of stable Higgs bundles
543: % is diffeomorphic to the moduli space of connections on $\Sigma$ of constant central
544: % curvature. Taking this constant to be zero, we obtain the (singular) moduli space
545: % of representations of the fundamental group of $\Sigma$.
546: % This correspondence between Higgs bundles, flat connections, and representations
547: % of the fundamental group is known as {\em nonabelian Hodge theory}, and has
548: % been studied extensively by other authors.
549:
550: \vspace{.2cm}
551: Consider the action of the multiplicative group
552: $\gm$ on $\cot$ given by scalar multiplication of the fibers.
553: The action of the compact subgroup $\so\subs\gm$ is hamiltonian
554: with respect to the first symplectic form $\omega_1$, and descends
555: to a circle action on $\M$ with moment map $\Phi:\M\to\R$, which is a Morse-Bott function.
556: The geometry and topology
557: associated with this action will be our main object of study.
558: In Chapter \ref{analogues} we give a detailed discussion of the construction of $\M$,
559: along with the action of $\gm$. In the case where $\Phi$ is proper,
560: we describe a reducible subvariety $\core\subs\M$ called the {\em core} of $\M$,
561: onto which $\M$ retracts $\so$-equivariantly.
562: The core $\core$ has $\X$
563: as one of its components, and if $\M$ is smooth, then $\core$ is equidimensional of dimension
564: $\dim\X = \half\dim\M$. In particular, the fundamental classes of the components
565: of $\core$ provide a natural basis for the top degree cohomology of $\M$. This fact is
566: exploited for hypertoric varieties in \cite{HS}, and for quiver varieties in various
567: papers of Nakajima. Building on \cite{HP1}, this thesis is the first
568: unified treatment of hyperk\"ahler analogues and their cores,
569: encompassing both hypertoric varieties and quiver varieties.
570:
571: The geometry of the core of $\M$ will be one of two major themes that we consider.
572: In the case where $\M$ is a hypertoric variety, each of the components of the core
573: $\core$ is itself a toric variety (Lemma \ref{htcorecomp}), as first shown in \cite{BD}.
574: In section \ref{htcore}, we give an explicit description of the action of
575: $\gm$ and the gradient flow of $\Phi$ on each piece. The case of hyperpolygon spaces
576: is more interesting. The ordinary polygon space $\X$ is the moduli space of $n$-sided polygons
577: in $\R^3$ with fixed edge lengths. In Section \ref{modcore}, we show that the other
578: core components are smooth, and may themselves be interpreted as moduli spaces
579: of polygons in $\R^3$ satisfying certain conditions (Theorem \ref{us}).
580: Thus, for the special case of hyperpolygon spaces,
581: we have solved the following general problem:
582:
583: \begin{problem}
584: Given any moduli space $\X$ that can be constructed as a K\"ahler reduction
585: (or GIT quotient) of complex affine space, is it possible
586: to interpret the core of the hyperk\"ahler analogue
587: $\X$ as a union of moduli spaces corresponding to other,
588: related moduli problems?
589: \end{problem}
590:
591: Our second major theme will be the circle-equivariant cohomology ring of $\M$.
592: In Chapter \ref{ht} we compute the circle-equivariant
593: cohomology ring of a hypertoric variety, and as an application compute the
594: $\Zt = Gal(\C/\R)$ equivariant cohomology ring of the complement of a complex
595: hyperplane arrangement defined over $\R$. The purpose of Chapter \ref{abelianization} is
596: to extend to the hyperk\"ahler setting a theorem of Martin \cite{Ma}, which describes
597: how to compute the cohomology ring of a K\"ahler quotient $X\mod G$ in terms
598: of the cohomology ring of the abelian quotient $X\mod T$, where $T\subs G$ is a maximal
599: torus. The main technical difficulty arises from the fact that
600: Martin's theorem relies heavily on computing integrals, which is not possible
601: on the noncompact hyperk\"ahler analogues that we have defined.
602: Our approach is to make use of the localization theorem of \cite{AB,BV}, which
603: allows us to define an integration theory in the circle-equivariant cohomology
604: of $\so$-manifolds with compact, oriented fixed point set. This is perhaps the single
605: most important reason for considering the circle action on a hyperk\"ahler analogue.
606: In Section \ref{hp} we combine the results of Chapters \ref{ht} and \ref{abelianization}
607: to compute the equivariant cohomology ring of a hyperpolygon space, and of each
608: of its core components.
609:
610: Most of Chapter \ref{ht} (with the exception of Section \ref{cogs})
611: appeared first in \cite{HP1}, and Chapter \ref{abelianization}
612: is a reproduction of \cite{HP}. Chapter \ref{hpspaces}
613: is taken primarily from \cite{HP2}, with the exception of Section \ref{s1},
614: which comes from \cite{HP}.
615: \end{chapter}
616:
617: \begin{chapter}{Hyperk\"ahler analogues}\label{analogues}
618: Our plan for this chapter is to provide a unified approach to the constructions of
619: hypertoric varieties and quiver varieties, which are the two major classes of examples
620: of hyperk\"ahler analogues of familiar K\"ahler varieties that appear in the literature.
621: In Section \ref{reduction} we give the basic construction of the hyperk\"ahler analogue
622: $\M$ of a K\"ahler quotient $\X=\Cn\mod G$, and show that $\M$ may be understood
623: as a partial compactification of the cotangent bundle to $\X$
624: (Proposition \ref{denseopen}).
625: In Section \ref{circleaction}, we define a natural action of the group $\gm$
626: on $\M$, which is holomorphic with respect to one of the complex structures.
627: This action will be our main tool for studying the geometry of $\M$
628: in future chapters.
629: Some of this material appeared first in \cite[\S 1]{HP1}.
630:
631: \begin{section}{Hyperk\"ahler and holomorphic
632: symplectic reduction}\label{reduction}
633: A {\em hyperk\"ahler manifold} is a Riemannian manifold $(M,g)$ along with
634: three orthogonal, parallel complex structures, $J_1,J_2,J_3$,
635: satisfying the usual quaternionic relations.
636: These three complex structures allow us to define three symplectic forms
637: $$\omega_1(v,w) = g(J_{1}v, w), \hs\omega_2(v,w) = g(J_{2}v,w),
638: \hs\omega_3(v,w) = g(J_{3}v,w),$$ so that $(g, J_i, \omega_i)$ is a K\"ahler
639: structure on $M$ for $i=1,2,3$.
640: The complex-valued two-form $\omega_2 + i
641: \omega_3$ is closed, nondegenerate, and holomorphic
642: with respect to the complex structure $J_1$.
643: Any hyperk\"ahler manifold can therefore be considered as a {\em holomorphic
644: symplectic} manifold with complex structure $J_1$, real symplectic
645: form $\omega_{\R} := \omega_1$, and holomorphic symplectic form
646: $\omega_{\C} := \omega_2 + i\omega_3$.
647: This is the point of view that we will adopt.
648:
649: We will refer to an action of $G$ on a hyperk\"ahler manifold
650: $M$ as {\em hyperhamiltonian}
651: if it is hamiltonian with respect to $\omega_{\R}$ and holomorphic
652: hamiltonian with respect to $\omega_{\C}$, with $G$-equivariant
653: moment map
654: $$\mhk:=\mr\oplus\mc:M\to\gd\oplus\g_{\C}^*.$$
655: The following theorem describes the {\em hyperk\"ahler quotient}
656: construction, a quaternionic analogue of the K\"ahler quotient.
657:
658: \begin{theorem}\label{hklr}{\em\cite{HKLR}}
659: Let $M$ be a hyperk\"ahler manifold equipped with a hyperhamiltonian action of a
660: compact Lie group $G$, with moment maps $\mu_1, \mu_2, \mu_3$.
661: Suppose $\xi = \xi_{\R}\oplus \xi_{\C}$ is a
662: central regular value of $\mhk$, with $G$ acting freely
663: on $\mhk^{-1}(\xi)$. Then there is a unique
664: hyperk\"ahler structure
665: on the hyperk\"ahler quotient
666: $\M = M \mmod_{\!\!\xi}G := \mhk^{-1}(\xi)/G$,
667: with associated symplectic and holomorphic symplectic forms
668: $\omega^{\xi}_{\R}$ and $\omega^{\xi}_{\C}$, such that
669: $\omega^{\xi}_{\R}$ and $\omega^{\xi}_{\C}$ pull back to the restrictions
670: of $\omega_{\R}$ and $\omega_{\C}$ to $\mhk^{-1}(\xi)$.
671: \end{theorem}
672:
673: For a general regular value $\xi$, the action of $G$ on
674: $\mhk^{-1}(\xi)$ will not be free, but only locally free.
675: To deal with this situation we must introduce the notion
676: of a hyperk\"ahler orbifold.
677:
678: An {\em orbifold} is a topological space $M$ locally modeled
679: on finite quotients of euclidean space.
680: More precisely, $M$ is a Hausdorff topological space,
681: equipped with an {\em atlas} $\mathcal U$ {\em of uniformizing charts}.
682: This consists of a collection of quadruples
683: $(\tilde U, \Gamma, U, \phi)$, where $\tilde U$ is an open subset
684: of euclidean space, $\Gamma$ is a finite group acting on $\tilde U$ and
685: fixing a set of codimension at least $2$,
686: $U$ is an open subset of $M$, and $\phi$ is a homeomorphism from
687: $\tilde U/\Gamma$ to $U$. The sets $U$ are required to cover $M$,
688: and the quadruples must satisfy a list of compatibility
689: conditions, as in \cite{LT}.
690:
691: Given a point $p\in M$, the {\em orbifold group at $p$} is the isotropy
692: group $\Gamma_p\subs\Gamma$ of a point
693: $\tilde p \in \phi^{-1}(p)\subs\tilde U$
694: for any quadruple $(\tilde U, \Gamma, U, \phi)$ such that $U$ contains $p$.
695: The orbifold tangent space $T_pM = T_{\tilde p}\tilde U_{\tilde p}$
696: should be thought of not as a vector space,
697: but rather as a representation of $\Gamma_p$ (see Proposition \ref{coreprops}).
698: A differential form on an orbifold may be thought of as a collection
699: of $\Gamma$-invariant differential forms on the open sets $\tilde U$,
700: subject to certain compatibility conditions. We may define riemannian metrics,
701: complex structures, vector bundles, K\"ahler structures, and hyperk\"ahler
702: structures on orbifolds in a similar manner.
703:
704: \begin{example}
705: Let $Z$ be a smooth manifold, and let $G$ be a compact Lie group
706: acting locally freely on $Z$. Then $Z/G$ inherits the structure
707: of an orbifold. The orbifold group of an orbit of a point $z\in Z$ is simply
708: the stabilizer $G_z\subs G$.
709: Any $G$-invariant tensor on $Z$ descends to an orbifold
710: tensor on $Z/G$. Any $G$-equivariant vector bundle on $Z$ descends
711: to a vector bundle on $Z/G$.
712: {\em All of the orbifolds that we consider will be of this form}
713: (and it is not known whether any other examples exist).
714: \end{example}
715:
716: These definitions allow for a straightforward extension of
717: Theorem \ref{hklr} to the case where $\xi$ is an arbitrary regular value
718: of $\mhk$. This implies, by the moment map condition, that $G$
719: acts locally freely on $\mhk^{-1}(0)$, and that the quotient
720: $\mhk^{-1}(0)/G$ inherits the structure of a hyperk\"ahler orbifold.
721:
722: Orbifolds are in many ways just as nice as manifolds; for example,
723: it is possible to adapt Morse theory to the orbifold case, as in \cite{LT},
724: which we will use in the next section. When we say that a certain
725: K\"ahler or hyperk\"ahler quotient is an orbifold, we wish to express
726: the opinion that it is relatively well behaved, rather than
727: the opinion that it is nasty and singular. For this reason, we will
728: use the positively connoted adjective $\Q$-{\em smooth}
729: to refer to orbifolds.
730:
731: We now specialize to the case where $M = \cot$, and the action
732: of $G$ on $\cot$ is induced from a linear action of $G$ on $\Cn$
733: with moment map $\mu:\Cn \to \gd$.
734: Choose an identification of $\Hn$ with $\cot$
735: such that the complex structure $J_1$ on $\Hn$
736: given by right multiplication by $i$
737: corresponds to the natural complex structure on $\cot$.
738: Then $\cot$ inherits a hyperk\"ahler structure.
739: The real symplectic form $\omega_{\R}$ is given by
740: adding the pullbacks of the standard forms on $\Cn$ and $(\Cn)^*$,
741: and the holomorphic symplectic form $\omega_{\C} = d\eta$, where $\eta$ is the canonical
742: holomorphic 1-form on $\cot$.
743:
744: We note that $G$ acts $\H$-linearly on $\cot\cong\Hn$
745: (where $n\times n$ matrices
746: act on the left on the space of column vectors $\Hn$, and scalar multiplication by $\H$
747: is on the right).
748: This action is hyperhamiltonian
749: with moment map $\mhk=\mc\oplus\mr$,
750: where
751: $$\mr(z,w) = \mu(z)-\mu(w)\hspace{10pt}\text{and}
752: \hspace{10pt}\mc(z,w)(v) = w(\hat{v}_z)$$
753: for $w\in T^*_z\Cn,\hs v\in\g_{\C}$, and $\hat{v}_z$ the element
754: of $T_z\Cn$ induced by $v$.
755: Given a central element $\a\in\gd$,
756: we refer to the hyperk\"ahler quotient $$\M = \cot \mmod_{\!\!(\a,0)} G$$
757: as the {\em hyperk\"ahler analogue} of the corresponding K\"ahler quotient
758: $$\X=\Cn\mod_{\!\!\a} G := \mu^{-1}(\a)/G.$$
759: In future sections we will often fix a parameter $\a$ and drop it from the notation.
760:
761: At times it will be convenient to think of K\"ahler quotients
762: in terms of geometric invariant theory, as follows.
763: Let $G_{\C}$ be an algebraic group acting on an affine variety $V$,
764: and let $\chi:G_{\C}\to\gm$ be a character of $G_{\C}$.
765: This defines a lift of the action of $G_{\C}$ to the trivial line
766: bundle $V\times \C$ by the formula
767: $$g\cdot(v,z) = (g\cdot v, \chi(g)^{-1}z).$$
768: The {\em semistable locus} $V^{s}$ with respect to
769: $\chi$ is defined to be the set
770: of points $v\in V$ such that for $z\neq 0$, the closure
771: of the orbit $G_{\C}(v,z)\subs V\times\C$ is disjoint from the
772: zero section $V\times\{0\}$ (see \cite{MFK} or \cite{N4}).
773: The {\em geometric invariant theory} (GIT) quotient $V\mod_{\!\!\chi}G_{\C}$
774: of $V$ by $G_{\C}$ at $\chi$ is an algebraic variety with underlying
775: space $V^{ss}\!/\!\!\sim$, where $v\sim w$
776: if and only if the closures $\overline{G_{\C}v}$
777: and $\overline{G_{\C}w}$ intersect in $V^{ss}$.
778: The {\em stable locus} $V^{st}$ with respect to $\chi$ is the set of points
779: $v\in V^{ss}$ such that the $G_{\C}$ orbit through $v$ is closed in
780: $V^{ss}$. Clearly the geometric quotient $V^{st}\!/G_{\C}$ is an open set
781: inside of the categorical quotient $V\mod_{\!\!\chi} G_{\C}$.
782: The following theorem is due to Kirwan in the projective case \cite{Ki};
783: our formulation of it is taken from \cite[\S 3]{N4}
784: and \cite[\S 8]{MFK}.
785:
786: \begin{theorem}\label{hkhs}
787: Let $G$ be a compact Lie group acting linearly on a complex vector
788: space $V$ with moment map $\mu:V\to\gd$. Let $G_{\C}$ be the complexification
789: of $G$, with its induced action on $V$. Let $\chi$ be a character of
790: $G$, and let $d\chi$ be the associated element of 5 center of $\gd$.
791: Then $v\in V^{ss}$ if and only if $G_{\C}v\cap\mu^{-1}(d\chi)\neq\emptyset$,
792: and the inclusion $\mu^{-1}(d\chi)\subs V^{ss}$ induces a homeomorphism
793: from $V\mod_{\! d\chi} G$ to $V\mod_{\!\chi} G_{\C}$.
794: Furthermore, $d\chi$ is a regular
795: value of $\mu$ if and only if $V^{ss}=V^{st}$.
796: \end{theorem}
797:
798: Given a regular value $\a\in\gd$, Theorem \ref{hkhs} tells
799: us that we may interpret $V\mod_{\!\!\a} G$ as a GIT quotient
800: only in the case where $\a$ comes from a character of $G$.
801: We note, however, that the stability and semistability conditions
802: are unchanged when $\chi$ is replaced by a high power of itself,
803: hence we may apply Theorem \ref{hkhs} whenever some multiple
804: of $\a$ comes from a character. Furthermore, the GIT stability condition
805: is locally constant as a function of $\chi$. Hence for any central
806: regular value $\a\in\gd$,
807: we may perturb $\a$ to a ``rational'' point, thereby
808: interpret $V\mod_{\!\!\a}G$ as a GIT quotient.
809: Accordingly, we will call an element of $V$ stable with respect
810: to $\a\in\gd$ if and only if it is stable with respect to $\chi$
811: for all $\chi$ such that $d\chi$ is close to a multiple of $\a$,
812: and we will write $V\mod_{\!\!\a}G_{\C} = V^{ss}\!/\!\!\sim$.
813:
814: We may also use this theorem to formulate the hyperk\"ahler quotient
815: construction purely in terms of algebraic geometry.
816: Theorem \ref{hkhs} says that, for $\a$ a regular value of $\mur$,
817: $$\cot\mod_{\!\!\a}G\,\cong\,\cot\mod_{\!\!\a}G_{\C}\,\cong\,\(\cot\)^{st}\!/G_{\C}.$$
818: Since $\muc:\cot\to\gd_{\C}$ is equivariant, we may take its vanishing locus
819: on both sides of the above equation, and we obtain the identity
820: $$\cot\mmod_{\!\!(\a,0)}G\,\cong\,\muc^{-1}(0)\mod_{\!\!\a}G_{\C}
821: \,\cong\,\muc^{-1}(0)^{st}/G_{\C}.$$
822:
823: The following proposition is proven for the case
824: where $G$ is a torus in \cite[7.1]{BD}.
825:
826: \begin{proposition}\label{denseopen}
827: Suppose that $\a$ and $(\a,0)$ are regular values for $\mu$ and $\mhk$,
828: respectively.
829: The cotangent bundle $T^*\X$ is isomorphic to an open subset of $\M$,
830: and is dense if it is nonempty.
831: \end{proposition}
832:
833: \begin{proof}
834: Let $Y = \{(z,w)\in\mc^{-1}(0)^{st}\mid
835: z \in (\Cn)^{st}\}$,
836: where we ask $z$ to be semistable with respect to $\a$
837: for the action of
838: $G_{\C}$ on $\Cn$, so that $\X\cong (\Cn)^{st}/G_{\C}$.
839: Let $[z]$ denote the element of $\X$ represented by $z$.
840: The tangent space $T_{[z]}\X$ is equal to the quotient of $T_{z}\Cn$
841: by the tangent space to the $G_{\C}$ orbit through $z$,
842: hence
843: $$T^*_{[z]}\X\cong\{w\in T^*_{z}\Cn\mid w(\hat{v}_z)=0\text{ for all }
844: v\in\g_{\C}\} = \{w\in(\Cn)^*\mid\mc(z,w)=0\}.$$
845: Then $$T^*\X \cong \{(z,w)\mid z\in(\Cn)^{st}\text{ and }
846: \mc(z,w)=0\}/G_{\C} = Y/G_{\C}.$$
847: By the definition of semistability, $Y$ is an open subset of $\mc^{-1}(0)$,
848: and is dense if nonempty. This completes the proof.
849: \end{proof}
850:
851: \vspace{-\baselineskip}
852: \begin{remark}
853: We may significantly generalize the construction of hyperk\"ahler analogues
854: as follows. Replace $\Cn$ by a smooth complex variety $X$, equipped with an
855: action of an algebraic group $G_{\C}$, an ample line bundle $L$, and a lift
856: of the action to $L$. Then the cotangent bundle $T^*X$ is holomorphic symplectic,
857: and carries a natural holomorphic hamiltonian action of $G_{\C}$, along with a lift
858: of this action to the pullback of $L$. We may then define the holomorphic symplectic
859: analogue of the GIT quotient $\X = X\mod G_{\C}$ to be the GIT quotient of the
860: zero level of the holomorphic moment map in $T^*X$, where the semistable
861: sets are defined by the action of $G_{\C}$ on $L$. Theorem \ref{hkhs} tells us
862: that this agrees with our construction if $X=\Cn$, and Proposition \ref{denseopen}
863: generalizes to say that the holomorphic symplectic analogue of $\X$ is a partial
864: compactification of its cotangent bundle.
865:
866: The reason for relegating this definition to a remark is that when $X$
867: is not equal to $\Cn$, its cotangent bundle $T^*X$ may not be the best holomorphic
868: symplectic manifold with which to replace it. For example, if $X$ is itself a
869: K\"ahler quotient, then the holomorphic symplectic analogue of $X$ modulo the trivial
870: group, in the sense of the previous paragraph, would simply be the cotangent bundle
871: to $X$. But this would (usually) not agree with the hyperk\"ahler
872: analogue of $X$.
873: \end{remark}
874:
875: \end{section}
876:
877: \begin{section}{The $\bf{\gm}$ action and the core}\label{circleaction}
878: Consider the action of $\gm$ on $T^*\Cn$ given by
879: scalar multiplication on the fibers, that is $\tau\cdot(z,w) = (z,\tau w)$.
880: The holomorphic moment map $\muc:\cot\to\gdc$ is $\gm$-equivariant
881: with respect to the scalar action on $\gdc$,
882: hence $\gm$ acts on $\muc^{-1}(0)$. Linearity of the action of $G$
883: on $\Cn$ implies that the actions of $\GC$ and $\gm$ on $\cot$ commute,
884: therefore we obtain a $J_1$-holomorphic action of $\gm$ on
885: $\Ma = \muc^{-1}(0)\mod\GC$.
886: Note that the $\gm$ action does {\em not} preserve the holomorphic symplectic form
887: or the hyperk\"ahler structure on $\Ma$;
888: rather we have $\t^*\omega_{\C}=\t\omega_{\C}$ for $\t\in\gm$.
889:
890: If $\Ma$ is $\Q$-smooth, then the action of the compact subgroup
891: $\so\subs\gm$ is hamiltonian with respect to the real symplectic
892: structure $\omega_{\R}$, with moment map
893: $\Phi[z,w]_{\R} = \half |w|^2$.
894: This map is an orbifold Morse-Bott function\footnote{For a detailed discussion
895: of hamiltonian circle actions and Morse theory on orbifolds,
896: see \cite{LT}.} with image
897: contained in the non-negative real numbers, and
898: $\Phi^{-1}(0)=\Xa\subs\Ma$.
899:
900: \begin{proposition}\label{proper}
901: If the original moment map $\mu:\Cn\to\gd$ is proper,
902: then so is $\Phi:\Ma\to\R$.
903: \end{proposition}
904:
905: \begin{proof}
906: We must show that $\Phi^{-1}[0,R]$ is compact
907: for any $R\in\R$. We have
908: $$\Phi^{-1}[0,R] = \{(z,w)\mid \mr(z,w)=\a,\hs
909: \mc(z,w)=0,\hs\Phi(z,w)\leq R\}\big/G$$ and $G$ is compact,
910: hence it is sufficient to show that
911: $\{(z,w)\mid \mr(z,w)=\a,\hs\Phi(z,w)\leq R\}$
912: is compact. Since $\mr(z,w)=\mu(z)-\mu(w)$, this set is a closed subset
913: of $$\mu^{-1}\left\{\a+\mu(w) \,\Bigg{|}\, \half |w|^2\leq R\right\}
914: \times\left\{w \,\Bigg{|}\, \half |w|^2 \leq R\right\},$$
915: which is compact by the properness of $\mu$.
916: \end{proof}
917:
918: \vspace{-\baselineskip}
919: \begin{remark}\label{tc}
920: In the case where $G$ is abelian and $\Xa$ is a nonempty toric
921: variety, properness of $\mu$ (and therefore of $\Phi$) is equivalent
922: to compactness of $\Xa$.
923: \end{remark}
924: % \end{section}
925:
926: % \begin{section}{The core}
927: Suppose that $\Ma$ is $\Q$-smooth and $\Phi$ is proper.
928: We define the {\em core} $\core\subs\Ma$ to be the union
929: of those $\gm$ orbits whose closures are compact.
930: Properness of $\Phi$ implies that $\Ld_{\t\to 0}\t\cdot p$ exists for all $p\in \M$,
931: hence we may write
932: $$\core = \{p\in\Ma\mid\Ld_{\t\to\infty}\t\cdot p\text{ exists}\}.$$
933: For $F$ a connected component of $\Mso=\Macs$,
934: let $U_F$ be the closure of the set
935: of points $p\in \M$ such that $\Ld_{\t\to\infty}\t\cdot p\in F$.
936: In Morse-theoretic language,
937: $U(F)$ is the closure of the unstable orbifold of the critical set $F$.
938: We may then write $\core$ as a finite union of irreducible, compact
939: varieties as follows: $$\core = \bigcup_{F\subs\Macs}U_F.$$
940:
941: \begin{proposition}\label{coreprops}
942: The core of $\Ma$ has the following properties:
943: \begin{enumerate}
944: \item $\core$ is an $\so$-equivariant deformation retract of $\Ma$
945: \item $U_F$ is isotropic with respect to $\omega_{\C}$
946: \item If $\Ma$ is smooth at $F$, then $\dim U_F = \half\dim\Ma$.
947: \end{enumerate}
948: \end{proposition}
949:
950: \begin{proof}
951: Let $f:\Ma\to [0,1]$ be a smooth, $\so$-invariant function with $f^{-1}(0)=\core$.
952: For all $p\in \M$ and $t\geq 0$, let $\rho_t(p) = e^{f(p)t}\cdot p$.
953: This defines a smooth family of $\so$-equivariant maps $\rho_t:\Ma\to\Ma$,
954: fixing $\core$, with $\rho_0 = \id$.
955: The limit $\rho_{\infty} = \Ld_{t\to\infty}\rho_t$ is a well-defined
956: smooth map from $\M$ to $\core$, hence (1) is proved.
957:
958: Suppose that $\Ma$ is smooth at $F$ and consider a point $p\in F$.
959: Since $p$ is a fixed point, $\so$ acts on $T_p\Ma$,
960: and we may write $$T_p\Ma = \Od_{s\in\Z}H_s,$$
961: where $H_s$ is the $s$ weight space for the circle action.
962: Since $\t^*\omega_{\C}=\t\omega_{\C}$ and $\omega_{\C}$ is a nondegenerate
963: bilinear form on $T_p\Ma$, $\omega_{\C}$ restricts to a perfect pairing
964: $H_s\times H_{1-s}\to\C$ for all $s\in\Z$.
965: In particular, $$T_pU_F = \Od_{s\leq 0}H_s$$ is a maximal isotropic subspace
966: of $T_p\Ma$, thus proving (2) and (3).
967:
968: Now suppose that $\Ma$ is only $\Q$-smooth at $F$, and let $\Gap$ be the
969: orbifold group at $p$. A circle action on the orbifold tangent
970: space $T_p\M$ is an action of a group $\Gapt$,
971: where $\Gapt$ is an extension of $\so$ by $\Gap$.
972: Let $\Gapto$ be the connected
973: component of the identity in $\Gapt$. Then $\Gapto\,$ is itself isomorphic
974: to a circle, and maps to the original circle $\so$ with some degree $d\geq 1$.
975: We now decompose $T_p\Ma = \bigoplus H_s$ into $\Gapto$ weight spaces.
976: Again $\omega_{\C}$ is nondegenerate on $T_p\Ma$, but now
977: $\hat{\t}^*\omega_{\C} = \hat{\t}^d\omega_{\C}$
978: for $\hat{\t}\in\Gapto\cong\so$, hence $\omega_{\C}$ restricts to a perfect
979: pairing $H_s\times H_{d-s}\to\C$ for all $s\in\Z$.
980: It follows that $T_pU_F = \bigoplus_{s\leq 0}H_s$ is isotropic (though not necessarily
981: maximally isotropic\footnote{See Example \ref{orbi}.})
982: with respect to $\omega_{\C}$. This completes the proof
983: of (2) in the orbifold case.
984: \end{proof}
985:
986: \vspace{-\baselineskip}
987: \begin{remark}\label{cpt}
988: Proposition \ref{coreprops} provides a new way to understand Proposition \ref{denseopen}
989: in the case where $\M$ is $\Q$-smooth and $\Phi$ is proper.
990: The K\"ahler quotient $\Xa$ is an $\omega_{\C}$-lagrangian
991: suborbifold of $\Ma$, hence $\omega_{\C}$ identifies the normal
992: bundle to $\Xa$ in $\Ma$ with the cotangent bundle of $\Xa$.
993: The Proposition \ref{denseopen} follows from the fact that the normal bundle to
994: $\Xa$ in $\Ma$ can be identified with the dense open set of points in $\Ma$
995: that flow down to $\Xa=\Phi^{-1}(0)$ along the gradient of $\Phi$.
996: This also demonstrates that the $\gm$ action on $\M$ restricts to scalar multiplication
997: on the fibers of $T^*\Xa$.
998: \end{remark}
999:
1000: Given a space $M$ equipped with the action of a group $G$, we say that $M$
1001: is {\em equivariantly formal} if the equivariant cohomology ring
1002: $H^*_G(M)$ is a free module over $H^*_G(pt)$.
1003: We end the section with the statement of a fundamental theorem
1004: which we will use repeatedly throughout the paper. Theorem \ref{formality}
1005: is proven in the compact case in \cite{Ki}, and the proof goes through
1006: in the noncompact case as long as $\Phi$ is proper
1007: (see, for example, \cite[\S 2.2]{H1} or \cite[4.2]{TW}).
1008:
1009: \begin{proposition}\label{formality}
1010: Let $M$ be a symplectic
1011: orbifold with a hamiltonian circle action
1012: such that the moment map $\Phi:M\to\R$ is proper and bounded below,
1013: and has finitely many critical values.
1014: Then $H^*_{\so}(M)$ is a free module over $H^*_{\so}(pt)$.
1015: Moreover, if the action of $\so$ commutes with the action of another
1016: torus $T$, then $H^*_{T\times\so}(M)$ is a free module
1017: over $H^*_{\so}(pt)$.
1018: \end{proposition}
1019: \end{section}
1020: \end{chapter}
1021:
1022: \begin{chapter}{Hypertoric varieties}\label{ht}
1023: In this chapter we give a detailed analysis of the construction described
1024: in Chapter \ref{analogues} in the special case where the group $G$ is abelian.
1025: In this case the K\"ahler quotient $\Xa$ is called a {\em toric variety}, and
1026: its hyperk\"ahler analogue $\Ma$ is called a {\em hypertoric variety}.\footnote{Also
1027: known as a {\em toric hyperk\"ahler variety}.}
1028: These latter spaces were first studied
1029: systematically in \cite{BD}; other references include \cite{K1}, \cite{HS}, and
1030: \cite{HP1}.
1031:
1032: Just as there is a strong relationship between the geometry of toric
1033: varieties and the combinatorics of polytopes (see, for example, \cite{St}),
1034: the geometry of hypertoric
1035: varieties interacts with the combinatorics of real hyperplane arrangements.
1036: Hausel and Sturmfels \cite{HS} gave an interpretation of the cohomology
1037: ring of a hypertoric variety as the Stanley-Reisner ring of the matroid
1038: associated to the corresponding arrangement of hyperplanes (Theorem \ref{htm}).
1039: In Section \ref{htcoh}
1040: we interpret the $\so$-equivariant cohomology ring as an invariant of the oriented
1041: matroid, a richer combinatorial structure that can be associated to a hyperplane
1042: arrangement that is defined over the real numbers (Theorem \ref{htsm}
1043: and Remark \ref{om}).
1044: This result is applied in Section \ref{ossection} to obtain a combinatorial
1045: presentation of the $\Zt$-equivariant cohomology ring of the complement
1046: of an arrangment of complex hyperplanes defined over $\R$, thus enhancing
1047: the classical result of Orlik and Solomon \cite{OS}.
1048: In Section \ref{cogs}, we use the cogenerator approach of \cite{HS}
1049: to explore an aspect of the relationship between the cohomology rings of toric and
1050: hypertoric varieties.
1051: Most of the material presented here, with the exception of
1052: the entirety of Section \ref{cogs}, has been taken from \cite{HP1} in a modified form.
1053:
1054: \begin{section}{Hypertoric varieties and hyperplane arrangements}\label{arrangements}
1055: Let $\tn$ and $\td$ be real vector spaces of dimensions $n$ and $d$, respectively,
1056: with integer lattices $\tnz\subs\tn$ and $\tdz\subs\td$.
1057: Let $\{x_1,\ldots,x_n\}$ be an integer basis for $\tnz$,
1058: and let $\{\bd_1,\ldots,\bd_n\}$ be the dual basis for the dual lattice
1059: $\tndz\subs\tnd$.
1060: Suppose given $n$ nonzero integer vectors
1061: $\{\a_1,\ldots,\a_n\}\subs\tdz$ that span $\td$ over the real numbers.\footnote{In
1062: each of \cite{BD,K1,HS,HP1}, some additional assumption is placed on the vectors
1063: $\{a_i\}$. Sometimes they are assumed to be primitive, and sometimes
1064: they are assumed to generate the lattice $\tdz$ over the integers. Here
1065: we make neither assumption.}
1066: Define $\pi:\tn\to\td$ by $\pi(x_i)=a_i$, and let $\tk$ be the kernel of $\pi$,
1067: so that we have an exact sequence
1068: $$0 \longrightarrow \tk \stackrel{\i}{\longrightarrow} \tn
1069: \stackrel{\pi}{\longrightarrow} \td\longrightarrow 0.$$
1070: % Because $\{a_i\}$ spans $\tdz$, this sequence
1071: % exponentiates to an exact sequence of compact tori
1072: This sequence exponentiates to an exact sequence of abelian groups
1073: $$0 \longrightarrow \Tk \stackrel{\i}{\longrightarrow} \Tn
1074: \stackrel{\pi}{\longrightarrow} \Td\longrightarrow 0,$$
1075: where $$\Tn=\tn/\tnz,\hs\hs\Td=\td/\tdz,\hs\hs\text{and }\Tk=\ker(\pi:\Tn\to\Td).$$
1076: Thus $\Tk$ is a compact abelian group with Lie algebra $\tk$,
1077: which is connected if and only if the vectors $\{a_i\}$ span the lattice
1078: $\tdz$ over the integers.
1079: It is clear that every closed subgroup of $\Tn$ arises in this way.
1080:
1081: The restriction to $\Tk$ of the standard action of $\Tn$ on $\cot$
1082: is hyperhamiltonian with hyperk\"ahler moment map
1083: $$\mr\oplus\mc:\cot\to\tkd\oplus(\tk_{\C})^*,$$ where
1084: $$\mur(z,w) = \i^* \left(\half \sum_{i=1}^n (|z_i|^2 - |w_i|^2) \bd_i
1085: \right) \hspace{15pt}
1086: \text{and}\hspace{15pt}\muc(z,w) = \i^*_{\C} \left(\sum_{i=1}^n (z_i w_i)\bd_i\right).$$
1087: Given an element $\a\in\tkd$ with lift $r = (r_1,\ldots,r_n)\in\tnd$,
1088: the K\"ahler quotient
1089: $$\Xa= \Cn\mod\Tk = \mu^{-1}(\a)/\Tk$$
1090: is called a {\em toric variety}, and its hyperk\"ahler
1091: analogue $$\Ma=\cot\mmod\Tk =
1092: \Big(\mur^{-1}(\a)\cap\muc^{-1}(0)\Big)\Big/\Tk$$
1093: is called a {\em hypertoric variety}.
1094: Both of these spaces admit an effective residual action of the torus $\Td=\Tn/\Tk$
1095: which is hamiltonian in the case of $\Xa$, and hyperhamiltonian in the case
1096: of $\Ma$, with hyperk\"ahler moment map
1097: \begin{eqnarray*}
1098: \barmur[z,w]_{\R}\oplus\barmuc[z,w]_{\R}
1099: &=& \half \sum_{i=1}^n (|z_i|^2 - |w_i|^2 - r_i)\,\bd_i
1100: \hs\oplus\hs\sum_{i=1}^n (z_i w_i)\,\bd_i\\
1101: &\in&\ker(\i^*)\oplus\ker(\i^*_{\C}) = \tdd\oplus(\td_{\C})^*.
1102: \end{eqnarray*}
1103: In fact, this property may be used to give intrinsic definitions
1104: of toric and hypertoric varieties in certain categories,
1105: as demonstrated by the following two theorems.
1106: % \footnote{
1107: % Theorem \ref{bielawski} is proven only in the smooth case, but presumably
1108: % a similar result holds for orbifolds.}
1109:
1110: \begin{theorem}{\em\cite{De,LT}}
1111: Any connected symplectic orbifold of real dimension $2d$ which admits
1112: an effective, hamiltonian $\Td$ action with proper moment map is
1113: $\Td$-equivariantly symplectomorphic to a toric variety.
1114: \end{theorem}
1115:
1116: \begin{theorem}\label{bielawski}{\em\cite{Bi}}
1117: Any complete, connected, hyperk\"ahler manifold of real dimension $4d$
1118: which admits an effective, hyperhamiltonian $\Td$ action is $\Td$-equivariantly
1119: diffeomorphic, and Taub-NUT deformation equivalent, to a hypertoric variety.
1120: \end{theorem}
1121:
1122: The data that were used to construct $\Xa$ and $\Ma$ consist of a collection of
1123: nonzero vectors $a_i\in\tdz$ and an element $\a\in\tkd$. It is convenient
1124: to encode in terms of an arrangement of affine hyperplanes in $\tdd$ with some additional
1125: structure.
1126: A {\em rational, weighted, cooriented, affine hyperplane}
1127: $H\subs\tdd$ is an affine hyperplane
1128: along with a choice of nonzero normal vector $a\in\tdz$. The word rational
1129: refers to integrality of $a$, and weighted means that $a$ is not required
1130: to be primitive.
1131: % Let $r = (r_1,\ldots,r_n)\in\tnd$ be a lift of $\a$ along $\i^*$, and
1132: Consider the rational, weighted, cooriented hyperplane
1133: $$H_i =\{v\in\tdd \mid v\cdot a_i + r_i = 0\}$$
1134: with normal vector $a_i\in\tdz$, along with the two half-spaces
1135: \begin{equation}\label{fg}
1136: F_i = \{v\in\tdd \mid v\cdot a_i + r_i \geq 0\}\hspace{10pt}\text{and}
1137: \hspace{10pt}G_i = \{v\in\tdd \mid v\cdot a_i + r_i \leq 0\}.
1138: \end{equation}
1139: Let $$\D = \bigcap_{i=1}^n F_i=\{v \mid v\cdot a_i+r_i\geq 0
1140: \text{ for all }i\leq n\}$$
1141: be the (possibly empty) weighted polyhedron in $\tdd$ defined by the
1142: weighted, cooriented,
1143: affine, hyperplane arrangement $\A = \{H_1,\ldots,H_n\}$.
1144: % The condition that $\{a_1,\ldots,a_n\}$
1145: % spans $\tdz$ over the natural numbers implies that $\D$ is bounded
1146: % for every value of $r\in\tnd$, which in turn
1147: % implies compactness of $\Xa$.
1148: Choosing a different lift $r'$ of $\a$ corresponds combinatorially to translating
1149: $\A$ inside of $\tdd$, and geometrically to shifting the K\"ahler and hyperk\"ahler
1150: moment maps for the residual $\Td$ actions by $r'-r \in\ker\i^*=\tdd$.
1151: Our picture-drawing convention will be to encode the coorientations of the hyperplanes
1152: by shading $\D$, as in Figure \ref{teepee}. In every example that we consider,
1153: all hyperplanes will have weight $1$; in other words we
1154: will choose the primitive integer
1155: normal vector inducing the indicated coorientation.
1156:
1157: \begin{figure}[h]
1158: \centerline{\epsfig{figure=htv2.eps, height=4cm}}
1159: \caption{A cooriented arrangement representing a toric variety of complex dimension $2$,
1160: or a hypertoric variety of complex dimension $4$,
1161: obtained from an action of $T^2$ on $\C^4$.}
1162: \label{teepee}
1163: \end{figure}
1164:
1165: We call the arrangement $\A$ {\em simple} if every subset of $m$ hyperplanes
1166: with nonempty intersection intersects in codimension $m$.
1167: We call $\A$ {\em smooth} if every collection of $d$ linearly independent vectors
1168: $\{a_{i_1},\ldots,a_{i_d}\}$ spans $\tdd$.
1169: An element $r\in\tnd$ or $\a\in\tkd$ will be called simple if
1170: the corresponding arrangement $\A$ is simple.
1171:
1172: \begin{theorem}\label{simple}{\em\cite[3.2,3.3]{BD}}
1173: The hypertoric variety $\Ma$ is $\Q$-smooth if and only if $\A$ is simple,
1174: and smooth if and only if $\A$ is smooth.
1175: \end{theorem}
1176:
1177: % indexed by subsets $A\subs\{1,\ldots,n\}$.
1178: % Let $$\bmr\oplus\bmc:\Ma\to\tdd\oplus(\td_{\C})^*$$ be the hyperk\"ahler moment map
1179: % for the action of $\Td$ on $\Ma$.
1180: % For each $A\subs\{1,\ldots,n\}$, let $$M_A = \bmr^{-1}(\Delta_A)\cap\bmc^{-1}(0).$$
1181: % The K\"ahler submanifold
1182: % $\left(M_A,\omega_{\R}|_{M_A}\right)$ of $\left(M,\omega_{\R}\right)$
1183: % is $d$-dimensional and invariant under the action of $\Td$,
1184: % and is therefore $\Td$-equivariantly isomorphic
1185: % to the K\"ahler toric variety determined by $\Delta_A$ \cite[6.5]{BD}.
1186: % We define the {\em core} $C$ and {\em extended core} $D$
1187: % of a hypertoric variety by setting $$C = \!\!\bigcup_{\Delta_A\text{ bounded}}M_A\hspace{17pt}\text{ and }
1188: % \hspace{17pt} D = \bigcup_{\text{all }A}{M_A} = \mc^{-1}(0) =
1189: % \{[z,w]\mid z_iw_i=0\text{ for all }i\}.$$
1190: % Bielawski and Dancer \cite{BD} show that
1191: % $C$ and $D$ are each $\Td$-equivariant
1192: % deformation retracts of $M$. See Corollary \ref{unstable} for a Morse theoretic proof.
1193:
1194: Let us pause to point out the different ways in which
1195: $\Xa$ and $\Ma$ depend on the
1196: arrangement $\A$. The toric variety $\Xa$ is in fact determined by the weighted
1197: polyhedron $\D$ \cite{LT},
1198: and is therefore oblivious to any hyperplane $H_i$ such that $\D$ is contained in the interior of $F_i$.
1199: Thus the toric variety corresponding to Figure \ref{teepee} is
1200: $\C P^2$, the toric variety associated to a triangle.
1201: This is not the case for $\Ma$; we will see, in fact, that the
1202: hypertoric variety of Figure \ref{teepee} is topologically distinct
1203: from the one that we would obtain by deleting the third hyperplane.
1204: For this reason, it is slightly
1205: misleading to call $\Ma$ the hyperk\"ahler analogue of $\Xa$;
1206: more precisely, it is the hyperk\"ahler analogue of
1207: {\em a given presentation} of $\Xa$ as a K\"ahler quotient of a
1208: complex vector space.
1209:
1210: Just as the toric variety $\Xa$ fails to retain all of the data of
1211: the arrangement $\A$, there is some data that goes unnoticed by the
1212: hypertoric variety $\Ma$, as evidenced by the two following results.
1213:
1214: \begin{lemma}\label{placement}
1215: % If $\a, \a'\in\tdd$ are both simple, then
1216: % $M_{\a}$ and $M_{\a'}$ are $\Td$-equivariantly diffeomorphic, and their
1217: % cohomology rings can be naturally identified.
1218: The hypertoric variety $\Ma$ is independent, up to $\Td$-equivariant
1219: diffeomorphism,\footnote{Bielawski and Dancer \cite{BD}
1220: prove a weaker version of this statement,
1221: involving the (nonequivariant) homeomorphism type of $\Ma$.}
1222: of the choice of a simple element $\a\in\tkd$.
1223: \end{lemma}
1224:
1225: \begin{lemma}\label{coorientation}
1226: The hypertoric variety $\Ma$ is independent, up to $\Td$-equivariant isometry,
1227: of the coorientations of the hyperplanes $\{H_i\}$.
1228: \end{lemma}
1229:
1230: \begin{proofp}
1231: The set of nonregular values for $\mur\oplus\muc$
1232: has codimension 3 inside of
1233: $\tkd \oplus (\tk_{\C})^*$ \cite{BD}, hence we may
1234: choose a path connecting the two regular values $(\alpha,0)$
1235: and $(\alpha',0)$ for any simple $\a,\a'\in\tkd$,
1236: and this path is unique up to homotopy.
1237: Since the moment map $\mur\oplus\muc$
1238: is not proper, we must take some care in showing that
1239: two fibers are diffeomorphic. To this end, we note that the
1240: norm-square function $\psi(z,w) = \|z\|^2 + \|w\|^2$ is $T^n$-invariant
1241: and proper on $\cot$.
1242: Let $(\cot)_{reg}$ denote the open submanifold of $\cot$ consisting of the
1243: preimages of the regular values of $\mur\oplus\muc$.
1244: By a direct computation, it is easy
1245: to see that the kernels of $d\psi$ and $d\mur\oplus d\muc$ intersect
1246: transversely at any point $p \in (\cot)_{reg}$. Using the
1247: $T^n$-invariant hyperk\"ahler metric on $\cot$,
1248: we define an Ehresmann connection
1249: on $(\cot)_{reg}$ with respect to $\mur\oplus\muc$
1250: such that the horizontal subspaces are contained in the kernel of $d\psi$.
1251: %Note that this connection can be extended to an Ehresmann
1252: %connection with respect to $\iota^{*} \bar{\mu}_{\R} \oplus
1253: %\iota_{\C}^{*} \bar{\mu}_{\C} \oplus \psi.$
1254:
1255: This connection allows us to lift a path connecting the two
1256: regular values to a horizontal vector field on its preimage in
1257: $(\cot)_{reg}$. Since the horizontal subspaces are tangent to
1258: the kernel of $d\psi$, the flow preserves level sets of $\psi$. Note that the
1259: function $$\mur\oplus\muc\oplus\psi: \cot \to \tkd \oplus (\tk_{\C})^{*} \oplus\R$$
1260: {\em is} proper. By a theorem of Ehresmann \cite[8.12]{BJ}, the
1261: properness of this map implies that the flow of this vector field exists for all time, and
1262: identifies the inverse image of $(\alpha,0)$ with that of
1263: $(\alpha',0)$. Since the metric, $\psi$, and $\mur\oplus\muc$ are
1264: all $T^n$-invariant, the Ehresmann connection is also $T^n$-invariant,
1265: therefore the diffeomorphism identifying the fibers
1266: is $T^n$-equivariant, and
1267: the reduced spaces are $T^d$-equivariantly diffeomorphic.
1268: \end{proofp}
1269:
1270: \vspace{-\baselineskip}
1271: \begin{proofc}
1272: It suffices to consider the case when we change the orientation of a
1273: single hyperplane within the arrangement. Changing the coorientation
1274: of a hyperplane $H_m$ is equivalent to defining a new map
1275: $\pi':\tn\to\td$, with $\pi'(x_i)=a_i$ for $i\neq m$,
1276: and $\pi'(x_m) = -a_m$.
1277: This map exponentiates to a map $\pi':\Tn\to\Td$, with $\ker(\pi')$
1278: conjugate to $\ker(\pi)$ inside of $\operatorname{GL}_n(\H)$
1279: (the group of quaternion-linear
1280: automorphisms of $\cot\cong\Hn$) by the element
1281: $(1,\ldots,1,j,1,\ldots,1)\in \operatorname{GL}_1(\H)^n\subs \operatorname{GL}_n(\H)$,
1282: where the $j$ appears in the $m^{\text{th}}$ slot.
1283: Hence the hyperk\"ahler quotient by $\ker(\pi')$ is isomorphic
1284: to the hyperk\"ahler quotient by $\ker(\pi)$.
1285: \end{proofc}
1286:
1287: \vspace{-\baselineskip}
1288: \begin{example}\label{same}
1289: The three cooriented arrangements of Figure 2 all specify
1290: the same hyperk\"ahler variety $\M$ up to equivariant
1291: diffeomorphism. The first has $\X\cong\widetilde{\C P^2}$
1292: (the blow-up of $\C P^2$ at a point),
1293: and the second and the third have $\X\cong\C P^2$.
1294: Note that if we reversed the coorientation of $H_3$ in Figure 2(a)
1295: or 2(c), then we would get a noncompact
1296: $\X\cong\widetilde{\C^2}$. If we reversed the coorientation of $H_3$ in
1297: Figure 2(b), then $\X$ would be empty,
1298: but the topology of $\Ma$ would not change.
1299:
1300: \begin{figure}[h]
1301: \centerline{\epsfig{figure=htv1.eps}}
1302: \caption{Three arrangements
1303: related by reversing coorientations and translating hyperplanes.}
1304: \label{triplet}
1305: \end{figure}
1306: \end{example}
1307:
1308: Our purpose is to study not just the geometry
1309: of $\Ma$, but the geometry of $\Ma$ along with the hamiltonian circle action
1310: defined in Section \ref{circleaction}.
1311: In order to define this action, we used the fact that we were reducing at a regular value
1312: of the form $(\a,0)\in\tdd\oplus(\td_{\C})^*$.
1313: Although the set of regular values of $\mur\oplus\muc$ is simply connected,
1314: the set of regular values of the form $(\a,0)$ is disconnected, therefore the
1315: diffeomorphism of Lemma \ref{placement} is not circle-equivariant.
1316: Furthermore, left multiplication by the diagonal matrix
1317: $(1,\ldots,1,j,1,\ldots,1) \in
1318: \operatorname{GL}_n(\H)$ is not an $\so$-equivariant
1319: automorphism of $\cot\cong\Hn$,
1320: therefore the topological type of $\Ma$ along with a $\so$ action
1321: may depend nontrivially
1322: on both the locations and the coorientations of the hyperplanes in $\A$.
1323: Indeed it must, because we can recover $\Xa$ from $\Ma$ by taking the minimum
1324: $\Phi^{-1}(0)$ of the moment map $\Phi:\Ma\to\R$, and we know that $\Xa$ depends in an essential way
1325: on the combinatorial type of the polyhedron $\D$.
1326: In this sense, the structure of a hypertoric variety $\Ma$
1327: along with a hamiltonian circle action may be regarded as
1328: the universal geometric object associated to $\A$ from which
1329: both $\Ma$ and $\Xa$ can be recovered.
1330: \end{section}
1331:
1332: \begin{section}{Geometry of the core}\label{htcore}
1333: In this section we give a combinatorial description of the fixed point set
1334: $\M^{\gm} = \Ma^{\so}$
1335: and the core $\core$ of a $\Q$-smooth hypertoric variety $\Ma$.
1336: We will assume that $\Phi$ is proper. (If $\D$ is nonempty, this is equivalent to asking
1337: that $\D$ be bounded, or that $\Xa$ be compact.)
1338: % The parameter $\a$ will be fixed throughout
1339: % that section, hence we will omit the subscript.
1340: First, we note that the holomorphic moment map $\barmuc:\Ma\to(\td_{\C})^*$
1341: is $\gm$-equivariant with respect to the scalar action on $(\td_{\C})^*$,
1342: hence both $\Ma^{\gm}$ and $\core$ will be contained in
1343: $$\E=\barmuc^{-1}(0)=\Big\{[z,w]\in
1344: \Ma\bigmid z_iw_i=0\text{ for all } i\Big\},$$
1345: which we call the {\em extended core} of $\Ma$.
1346: It is clear from the defining equations that the restriction of
1347: $\barmur$ from $\E$ to $\tdd$ is surjective.
1348: The extended core naturally breaks into components
1349: $$\EA = \Big\{[z,w]\in \Ma\bigmid w_i=0
1350: \text{ for all $i\in A$ and $z_i=0$ for all $i\in A^c$}\Big\},$$
1351: indexed by subsets $A\subs\otn$.
1352: When $A=\emptyset$, $\EA = \X \subs \Ma$. More generally,
1353: the variety $\EA\subs \Ma$ is a $d$-dimensional K\"ahler subvariety
1354: of $\Ma$ with an effective hamiltonian $\Td$-action, and is therefore itself a toric variety.
1355: (It is the K\"ahler quotient by $\Tk$ of an $n$-dimensional
1356: coordinate subspace of $\cot$,
1357: contained in $\muc^{-1}(0)$.)
1358: The hyperplanes $\{H_i\}$ divide $\tdd$ into a union of closed,
1359: (possibly empty) convex polyhedra $$\Delta_A =
1360: \bigcap_{i\in A}F_i\,\,\cap\,\,\bigcap_{i\in A^c}G_i.$$
1361:
1362: \begin{lemma}\label{sides}
1363: If $w_i=0$, then $\barmur[z,w]_{\R}\in F_i$.
1364: If $z_i=0$, then $\barmur[z,w]_{\R}\in G_i$.
1365: \end{lemma}
1366:
1367: \begin{proof}
1368: We have
1369: $$\barmur[z,w]_{\R}\cdot a_i + r_i = \mur(z,w)\cdot x_i
1370: =\half\Big(|z_i|^2 - |w_i|^2\Big),$$
1371: hence the statement follows from Equation~\eqref{fg}.
1372: \end{proof}
1373:
1374: \vspace{-\baselineskip}
1375: \begin{lemma}\label{htcorecomp}{\em\cite{BD}}
1376: The core component $\EA$ is isomorphic to the toric variety
1377: corresponding to the weighted polytope $\D_A$.
1378: \end{lemma}
1379:
1380: \begin{proof}
1381: Lemma \ref{sides} tells us that $\barmur(\EA)\subs\Delta_A$,
1382: and surjectivity of $\barmur|_{\E}$ says that this inclusion is an equality.
1383: The lemma then follows from the classification theorems of \cite{De,LT}.
1384: \end{proof}
1385:
1386: Although $\gm$ does not act on $\Ma$ as a subtorus of $\Td_{\C}$, we show
1387: below that when restricted to any single
1388: component $\EA$ of the extended core, $\gm$
1389: {\em does} act as a subtorus of $\Td_{\C}$, with the subtorus depending
1390: combinatorially on $A$. This will allow us to give a combinatorial
1391: analysis of the gradient flow of $\Phi$ on the extended core.
1392:
1393: Suppose that $[z,w]\in\EA$. Then for $\tau\in\gm$,
1394: $$\tau[z,w]=[z,\tau w] =
1395: [\tau_1 z_1,\ldots \tau_n z_n, \tau_1^{-1}w_1,\ldots \tau_n^{-1}w_n], \text{ where }
1396: \tau_i =
1397: \begin{cases}
1398: \tau^{-1} & \text{if }i\in A,\\
1399: 1 & \text{if }i\notin A.
1400: \end{cases}
1401: $$
1402: In other words, the $\gm$ action on $\EA$
1403: is given by the one dimensional subtorus
1404: $(\tau_1, \ldots, \tau_n)$
1405: of the original torus $T^n_{\gm}$,
1406: hence the moment map $\Phi|_{\EA}$ for the action of $\so\subs\gm$ is given by
1407: $$\Phi[z,w] = \left<\mr[z,w],\,\,\sum_{i\in A}a_i\right>.$$
1408: This formula allows us to compute the fixed points of the circle action.
1409: % Since $S^1$ acts freely on $(\td_{\C})^{*} \backslash \{0\}$
1410: % and $\mc: \M \to (\td_{\C})^{*}$ is $S^1$-equivariant,
1411: % we must have $M^{S^1} \subseteq \mc^{-1}(0) = D$.
1412: For any subset $B\subs\{1,\ldots,n\}$, let $\EA^B$ be the toric subvariety
1413: of $\EA$ defined by the conditions $z_i=w_i=0$ for all $i\in B$.
1414: Geometrically, $\EA^B$ is defined by the (possibly empty) intersection of the hyperplanes
1415: $\{H_i\mid i\in B\}$ with $\Delta_A$.
1416: % $\displaystyle\bigcap_{i\in B}H_i \cap \Delta_A$ of the polyhedron $\Delta_A$.
1417:
1418: \begin{proposition}
1419: The fixed point set of the action of $S^1$ on $\EA$
1420: is the union of those toric subvarieties $\EA^B$ such that
1421: $\sum_{i\in A}a_i\in \td_B:=\operatorname{Span}_{j\in B}a_j$.
1422: \end{proposition}
1423:
1424: \begin{proof}
1425: The moment map $\Phi|_{\EA^B}$ will be constant if and only if
1426: $\sum_{i\in A}a_i$ is perpendicular to
1427: the face $\Phi(\EA^B)$,
1428: i.e. if $\sum_{i\in A}a_i$ lies in the kernel of the projection
1429: $\td\onto\td/\td_B$.
1430: \end{proof}
1431:
1432: \vspace{-\baselineskip}
1433: \begin{corollary}
1434: Every vertex $v\in\tdd$ of the polyhedral complex $|\A|$ given by our arrangement
1435: is the image of an $S^1$-fixed point in $\M$.
1436: Every component of $\M^{S^1}$ maps to a face of $|\A|$.
1437: % has dimension less than or equal to $d$,
1438: % and the only component of dimension $d$ is $\E_{\emptyset}=X=\Phi^{-1}(0)$.
1439: \end{corollary}
1440:
1441: \begin{proposition}
1442: The core $\core$ of $\Ma$ is equal to the union of those
1443: subvarieties $\EA$ such that $\Delta_A$ is bounded.
1444: \end{proposition}
1445:
1446: \begin{proof}
1447: Because $\gm$ acts on $\EA$ as a subtorus of the complex torus $\Td_{\C}$,
1448: the set $$\{p\in\EA\mid\Ld_{\t\to\infty}\t\cdot p\text{ exists}\}$$
1449: is a (possibly reducible) toric subvariety of $\EA$, i.e. a union of subvarieties
1450: of the form $\EA^B$. Fix a subset $B\subs\otn$.
1451: The variety $\EA^B$ is stable under the $\gm$ action, hence
1452: if $\EA^B$ is compact, then $\EA^B\subs\core$. On the other hand if $\EA^B$ is noncompact,
1453: then properness of $\Phi$ precludes it from being part of the core,
1454: hence $$\core = \{p\in\E\mid\Phi(p)\text{ lies on a bounded face of }|\A|\}.$$
1455: By \cite[6.7]{HS}, the bounded complex of $|\A|$ has pure dimension $d$,
1456: and is therefore equal to the union of those polyhedra $\Delta_A$ that are bounded.
1457: \end{proof}
1458:
1459: % For any point $p\in \M^{S^1}$, the {\em stable orbifold}
1460: % $S(p)$ at $p$ is defined to be the set of $x\in \M$ such
1461: % that $x$ approaches $p$ when flowing along the vector field
1462: % $-\operatorname{grad}(\Phi)$,
1463: % and the {\em unstable orbifold} $U(p)$ at $p$ is defined to be the stable orbifold with respect
1464: % to the function $-\Phi$.
1465: % For any suborbifold $Y\subs \M^{S^1}$, the unstable orbifold $U(Y)$ at $Y$ is defined
1466: % to be the union of $U(y)$ for all $y\in Y$. In general, for $y\in Y$, we have the identity
1467: % $\dim_{\R} U(Y) + \dim_{\R} S(y) = 4d$.
1468: %
1469: % Let $Y\subs \M^{S^1}$ be a component of the fixed point set of $M$.
1470: % Let $v\in\tdd$ be a vertex in the polytope $\mr(Y)$, and let $y$
1471: % be the unique preimage of $v$ in $Y$.
1472: %
1473: % \begin{proposition}
1474: % The unstable orbifold $U(Y)$ is a complex suborbifold of complex
1475: % dimension at most $d$, contained in the core $C\subs \M$.
1476: % If $\A$ is smooth at $y$, then $\dim_{\C}U(Y)=d$, and the closure of $U(Y)$
1477: % is an irreducible component of $C$.
1478: % \end{proposition}
1479: %
1480: % \begin{proof}
1481: % For simplicity, we will assume that $v=\cap_{j=1}^dH_j$.
1482: % For all $l\in\otd$, let $b_l\in\td_{\Z}$ be the smallest integer vector
1483: % such that $\left<a_j,b_l\>=0$ for $j\neq l$ and $\left<a_l,b_l\> >0$.
1484: % Geometrically, $b_l$ is the primitive integer vector on the line
1485: % $\cap_{j\neq l}H_j$ pointing in the direction of $\Delta$.
1486: % Note that $M$ is smooth at the $T^d$-fixed point above $v$
1487: % if and only if $\left<a_l,b_l\>=1$ for all $l\in\otd$.
1488: % Let $R_l\subs\tdd$ be the ray emanating from $v$ in the direction of $b_l$,
1489: % and ending before it hits another vertex. Let $Q_l$ be the analogous ray in the opposite
1490: % direction.
1491: %
1492: % Let $\Delta_A$ be a region (not necessarily bounded) of the polyhedral
1493: % complex defined by $\A$ adjacent to $R_l$.
1494: % The preimage $\mr^{-1}(R_l)\cap D$ of $R_l$ in $D$ is a complex line,
1495: % and it is contained in the unstable
1496: % orbifold at $U(Y)$ if and only if $\left<b_l,\sum_{i\in A}a_i\right>\geq 0$.
1497: % If $\left<b_l,\sum_{i\in A}a_i\right> <0$, it is contained in the stable orbifold $S(y)$.
1498: % The preimage $\mr^{-1}(Q_l)\cap D$ of $Q_l$ in $D$ is also a complex line,
1499: % contained in the unstable
1500: % orbifold $U(Y)$ if and only if $\<-b_l,a_l+\sum_{i\in A}a_i\>\geq 0$,
1501: % and otherwise in $S(y)$.
1502: % Since $\<b_l,\sum_{i\in A}a_i\>+\<-b_l,a_l+\sum_{i\in A}a_i\>
1503: % =-\left<a_l,b_l\> <0$, at most one of these two directions can be unstable.
1504: % In the smooth case, $\left<a_l,b_l\>=1$ for all $l$, and exactly one of
1505: % the two directions is unstable.
1506: %
1507: % Consider the polytope $\Delta_v$ incident to $v$ and characterized by the property that its
1508: % edges at the vertex $v$ are exactly the unstable directions.
1509: % The toric variety $X_{\Delta_v}\subs D$ is contained
1510: % in the closure of $U(Y)$, and a dimension count tells us that this containment is an equality.
1511: % In the smooth case, $\Delta_v$ is $d$-dimensional, and $X_{\Delta_v}$ is a component of the core.
1512: % \end{proof}
1513: %
1514: % Note that, even in the smooth case, it is not necessarily the case that the
1515: % $R_l$ direction is stable and the $Q_l$ direction is unstable. See, for example,
1516: % the vertex $v = H_1\cap H_2$ in Figure 2(c).
1517:
1518: \vspace{-\baselineskip}
1519: \begin{corollary}\label{bij}
1520: There is a natural injection from the set of bounded regions
1521: $\{\Delta_A\mid A\in I\}$
1522: to the set of connected components of $\M^{\gm}$.
1523: If $\A$ is smooth, this map is a bijection.
1524: \end{corollary}
1525:
1526: \begin{proof}
1527: To each $A\in I$, we associate the fixed subvariety $\EA^B$ corresponding
1528: to the face of $\Delta_A$ on which the linear functional $\sum_{i\in A}a_i$
1529: is minimized, so that $\EA=U(\EA^B)$.
1530: Proposition \ref{coreprops}~(3) tells us that if $\A$ is smooth and
1531: $F$ is a component
1532: of $\Ma^{\gm}$, then we will have $U(F)=\EA$ for some $A\subs\otn$.
1533: \end{proof}
1534:
1535: \vspace{-\baselineskip}
1536: \begin{example}
1537: In Figure 3, representing a reduction of $T^*\C^5$ by $T^3$,
1538: we choose a metric on $(\mathfrak{t}^2)^*$ in order
1539: to draw the linear functional $\sum_{i\in A}a_i$ as a vector
1540: in each region $\Delta_A$. We see that $\M^{S^1}$ has three components,
1541: one of them $X\cong\widetilde{\C P}^2$, one of them a projective line with another
1542: $\widetilde{\C P}^2$ as its associated core component, and one of them a point with
1543: core component $\C P^2$.
1544:
1545: \begin{figure}[h]
1546: \centerline{\epsfig{figure=flow.eps}}
1547: \caption{The gradient flow of $\Phi:\Ma\to\R$.}
1548: \end{figure}
1549: \end{example}
1550:
1551: \begin{example}\label{orbi}
1552: The hypertoric variety represented by Figure 4 has a fixed point set with four
1553: connected components (three points and a $\C P^2$), but only three components in its core.
1554: This phenomenon can be blamed on the orbifold point $p$ represented by the
1555: intersection of $H_3$ and $H_4$,
1556: which has only a one-dimensional unstable orbifold (to its northwest).
1557: In other words, this example illustrates the necessity of the smoothness
1558: assumption to obtain a bijection in Corollary \ref{bij}.
1559:
1560: \begin{figure}[h]
1561: \centerline{\epsfig{figure=orbi.eps}}
1562: \caption{A singular example.}
1563: \end{figure}
1564: \end{example}
1565: \end{section}
1566:
1567: \begin{section}{Cohomology rings}\label{htcoh}
1568: In this section we compute the $\so$ and $\Tdso$-equivariant cohomology
1569: rings of of a $\Q$-smooth hypertoric variety $\Ma$,
1570: extending the computations of the ordinary and $\Td$-equivariant
1571: rings given in \cite{K1} and \cite{HS}.
1572: For the sake of simplicity, and with an eye toward the applications in Chapter
1573: \ref{abelianization}, we will restrict our attention
1574: to the case where $\Phi$ is proper (see Remark \ref{tc}).
1575: This assumption will be necessary for the application of
1576: Proposition \ref{formality} and the proof of Theorem \ref{tara}.
1577:
1578: \begin{remark}\label{qorz}
1579: Because we wish to treat the smooth and $\Q$-smooth cases simultaneously,
1580: we will work with cohomology over the rational numbers. We note, however,
1581: that Konno proves Theorem \ref{htm} over the integers in the smooth case,
1582: and therefore our Theorem \ref{htsm} holds over the integers when $\A$ is smooth.
1583: This fact will be significant in Section \ref{ossection}, when we will need to reduce our
1584: coefficients modulo 2.
1585: \end{remark}
1586:
1587: Consider the hyperk\"ahler Kirwan maps
1588: $$\ktd:H^*_{\Tn}(\cot)\to\htm\hspace{15pt}\text{and}
1589: \hspace{15pt}\k:H^*_{\Tk}(\cot)\to H^*(\Ma)$$
1590: induced by the $\Tn$-equivariant inclusion of $\mur^{-1}(\a)\cap\muc^{-1}(0)$
1591: into $\cot$. Because the vector space $\cot$ is
1592: $\Tn$-equivariantly contractible, we have
1593: $$H^*_{\Tn}(\cot)=\Sym\tnd\cong\Q[\bd_1,\ldots,\bd_n]$$ and
1594: $$H^*_{\Tk}(\cot)=\Sym\tkd\cong\Q[\bd_1,\ldots,\bd_n]/\ker(\i^*).$$
1595:
1596: \begin{theorem}\label{htm}\label{hm}{\em\cite{K1,HS}}
1597: The Kirwan maps $\ktd$ and $\k$ are surjective, and the kernels of both are
1598: generated by the elements
1599: $$\prod_{i\in S}\bd_i\hspace{15pt}\text{ for all $S\subs\otn$ such that }
1600: \bigcap_{i\in S} H_i = \emptyset.$$
1601: \end{theorem}
1602:
1603: \begin{remark}
1604: The kernel of $\ktd$ is precisely
1605: the Stanley-Reisner ideal of
1606: the matroid of linear dependencies among the vectors $\{a_i\}$ \cite{HS}
1607: (see Remark \ref{om}).
1608: \end{remark}
1609:
1610: The inclusion of $\mur^{-1}(\a)\cap\muc^{-1}(0)$
1611: into $\cot$ is also $\so$-equivariant, hence we may consider the
1612: analogous circle-equivariant Kirwan maps
1613: $$\ktdso:H^*_{\Tn\times\so}(\cot)\to\htsm\hspace{15pt}\text{and}
1614: \hspace{15pt}\k:H^*_{\Tk\times\so}(\cot)\to H^*_{\so}(\Ma),$$
1615: where $$H^*_{\Tn\times\so}(\cot)\cong\Q[\bd_1,\ldots,\bd_n,x]$$ and
1616: $$H^*_{\Tk}(\cot)\cong\Q[\bd_1,\ldots,\bd_n,x]/\ker(\i^*).$$
1617: The remainder of this section will be devoted to proving the following theorem.
1618:
1619: \begin{theorem}\label{htsm}\label{hsm}
1620: The circle-equivariant Kirwan maps $\ktdso$ and $\kso$ are surjective,
1621: and the kernels of both are generated by the elements
1622: \begin{eqnarray*}
1623: \prod_{i\in S_1}\bd_i\times\prod_{j\in S_2}(x-\bd_j)\hspace{15pt}
1624: &\text{for}&\!\!\text{all disjoint pairs $S_1,S_2\subs\otn$}\\
1625: &&\text{such that }
1626: \bigcap_{i\in S_1} G_i\,\,\cap\,\,\bigcap_{j\in S_2} F_j = \emptyset.
1627: \end{eqnarray*}
1628: \end{theorem}
1629:
1630: \begin{remark}\label{om}
1631: For all $i\in\otn$, let $b_i = a_i\oplus 0\in\td\oplus\R$,
1632: and let $b_0=0\oplus 1$. The
1633: {\em pointed matroid} associated to $\A$ is a combinatorial object that
1634: tells us which subsets of $\{b_0,\ldots,b_n\}$ are linearly dependent.
1635: By simplicity of $\A$, this is equivalent to knowing which subsets
1636: $S\subs\otn$ have the property that $$\bigcap_{i\in S} H_i=\emptyset,$$
1637: which is in turn
1638: equivalent to knowing the dependence relations among the vectors $\{a_i\}$.
1639: In particular, it does not depend on the relative positions of the hyperplanes,
1640: encoded by the parameter $\a\in\tkd$.
1641:
1642: The {\em pointed oriented matroid} associated to $\A$ encodes the data
1643: of which subsets of $\{\pm b_0,\ldots,\pm b_n\}$ are linearly dependent
1644: over the {\em positive} real numbers. This is equivalent to knowing
1645: which pairs of subsets $S_1,S_2\subs\otn$ have the property that
1646: $$\bigcap_{i\in S_1} G_i\,\,\cap\,\,\bigcap_{j\in S_2} F_j = \emptyset,$$
1647: which does indeed depend on $\a$.
1648: Hence Theorem \ref{htm} shows that $\htm$ is an invariant
1649: of the pointed matroid of $\A$, and Theorem \ref{htsm} demonstrates that $\htsm$
1650: is an invariant of the pointed oriented matroid of $\A$.
1651: For more on this perspective, see \cite{H3} and \cite{Pr}.
1652: \end{remark}
1653:
1654: % In order to make use of Theorem \ref{htm}, we must apply
1655: % the principle of {\em equivariant formality},
1656: % proven for compact manifolds in \cite{Ki}, which we adapt to our situation
1657: % as follows.
1658: %
1659: % \begin{proposition}\label{formality}
1660: % Let $M$ be a symplectic
1661: % orbifold, possibly noncompact but of finite topological type.
1662: % Suppose that $M$ admits a hamiltonian action of a torus $T\times S^1$,
1663: % such that the fixed point set $M^{\so}$ has an even-dimensional,
1664: % $T$-equivariant cell decomposition,
1665: % and the $S^1$-component $\Phi:M\to\R$ of the moment map
1666: % is proper and bounded below. Then
1667: % $H^*_{T\times S^1}(M)$ is a free module over $H^*_{T\times\so}(pt)$.
1668: % \end{proposition}
1669: %
1670: % \begin{proof}
1671: % Because $\Phi$ is a moment map, it is a Morse-Bott function
1672: % such that all of the critical suborbifolds and their normal bundles
1673: % carry almost complex structures.
1674: % Thus we get a Morse-Bott stratification
1675: % of $M$ into even-dimensional $T\times\so$-invariant suborbifolds.
1676: % This tells us, as in \cite[5.8]{Ki},
1677: % that the spectral sequence associated to the fibration
1678: % $M\hookto EG\times_G \M\to BG$ collapses, where $G=T\times\so$,
1679: % and we get the desired result.
1680: % \end{proof}
1681: Consider the following commuting square of maps, where
1682: $\phi$ and $\psi$ are each given by setting the image of
1683: $x\in\ H^*_{\so}(pt)\cong\Q[x]$ to zero.
1684:
1685: $$\begin{CD}
1686: H^*_{\Tn\times S^1}(\cot) @>\ktdso >> \htsm\\
1687: @V\phi VV @ VV\psi V\\
1688: H^*_{\Tn}(\cot) @>\ktd >> \htm\\
1689: \end{CD}$$
1690:
1691: \begin{lemma}\label{surjective}
1692: The equivariant Kirwan map $\ktdso$ is surjective.
1693: \end{lemma}
1694:
1695: \begin{proof}
1696: Suppose that $\gamma\in\htsm$ is a homogeneous class of minimal degree
1697: that is {\em not} in the image of $\ktdso$. By Theorem \ref{htm} $\ktd$ is surjective,
1698: hence we may choose a class $\eta\in\phi^{-1}\ktd^{-1}\psi(\gamma)$.
1699: Theorem \ref{formality} tells us that the kernel of $\psi$ is generated by $x$,
1700: hence $\ktdso(\eta)-\gamma = x\delta$ for some $\delta\in\htsm$.
1701: Then $\delta$ is a class of lower degree that is not in the image of $\ktdso$.
1702: \end{proof}
1703:
1704: \vspace{-\baselineskip}
1705: \begin{lemma}\label{enough}
1706: If $\mathcal{I}\subs\ker\ktdso$ and $\phi(\mathcal{I}) = \ker\ktd$,
1707: then $\mathcal{I}=\ker\ktdso$.
1708: \end{lemma}
1709:
1710: \begin{proof}
1711: Suppose that $a\in\ker\ktdso\smallsetminus\mathcal{I}$
1712: is a homogeneous class of minimal degree,
1713: and choose $b\in\mathcal{I}$ such that $\phi(a-b)=0$. Then $a-b = cx$ for some
1714: $c\in H^*_{\Tn\times S^1}(\cot)$. By Proposition \ref{formality},
1715: $cx\in\ker\ktdso\impl c\in\ker\ktdso$, hence $c\in\ker\ktdso\smallsetminus\mathcal{I}$
1716: is a class of lower degree than $a$.
1717: \end{proof}
1718:
1719: \vspace{-\baselineskip}
1720: \begin{proofhtsm}
1721: For any element $$h\in H^2_{T^n\times S^1}(\cot)\cong\Q\{\bd_1,\ldots,\bd_n,x\},$$
1722: let $\tilde{L}_h = \cot\times\C_{h}$
1723: be the $\Tn\times S^1$-equivariant line bundle on $\cot$
1724: with equivariant Euler class $h$.
1725: Let $$L_h = \tilde{L}_h\Big{|}_{\mur^{-1}(\a)\cap\muc^{-1}(0)}\Big/\Tk$$
1726: be the quotient $\Td\times S^1$-equivariant line bundle on $\Ma$.
1727: We will write $\tilde L_i = \tilde L_{\bd_i}$ and
1728: $\tilde K = \tilde L_x$,
1729: with quotients $L_i$ and $K$.
1730: Since the $\Tdso$-equivariant Euler class $e(L_i)$
1731: is the image of $\bd_i$ under the hyperk\"ahler Kirwan map
1732: $H^*_{T^n\times S^1}(\cot)\to \htsm$, we will abuse notation and
1733: denote it by $\bd_i$. Similarly, we will denote $e(K)$ by $x$.
1734: Lemma \ref{surjective} tells us that
1735: $\htsm$ is generated by $\bd_1,\ldots,\bd_n,x$.
1736:
1737: Consider the $\Tn\times S^1$-equivariant section $\tilde s_i$ of $\tilde{L}_i$
1738: given by the function $\tilde s_i(z,w)=z_i$.
1739: This descends to a $\Tdso$-equivariant section $s_i$ of $L_i$
1740: with zero-set $$Z_i
1741: := \{[z,w]\in \Ma\mid z_i=0\}.$$
1742: Similarly, the function $\tilde t_i(z,w)=w_i$ defines a $\Tdso$-equivariant
1743: section of $L_i^*\otimes K$ with zero set $$W_i := \{[z,w]\in \Ma\mid w_i=0\}.$$
1744: Thus the divisor $Z_i$ represents the cohomology class $\bd_i$,
1745: and $W_i$ represents $x-\bd_i$.
1746: Note, that by Lemma \ref{sides},
1747: we have $\mr(Z_i) \subs G_i$ and $\mr(W_i) \subs F_i$ for all $i\in\otn$.
1748:
1749: Let $S_1$ and $S_2$ be a pair of subsets of $\{1,\ldots n\}$ such that
1750: $\big(\cap_{i\in S_1}G_i\big)\cap
1751: \big(\cap_{j\in S_2}F_j\big) = \emptyset,$ and hence
1752: $$\big(\cap_{i\in S_1}Z_i\big)\cap
1753: \big(\cap_{j\in S_2}W_j\big)\subs\mr^{-1}\bigg(
1754: \big(\cap_{i\in S_1}G_i\big)\cap
1755: \big(\cap_{j\in S_2}F_j\big)\bigg) = \emptyset.$$
1756: Now consider the vector bundle
1757: $E=\Od_{i\in S_1}L_i\,\,\oplus\,\,\Od_{j\in S_2}L_j^*\otimes K$
1758: with equivariant Euler class
1759: $$e(E)=\prod_{i\in S_1}\bd_i\times\prod_{j\in S_2}(x-\bd_j).$$
1760: The section
1761: $\left(\oplus_{i\in S_1}s_i\right)\oplus\left(\oplus_{i\in S_2}t_i\right)$
1762: is a nonvanishing equivariant
1763: global section of $E$,
1764: hence $e(E)$ is trivial in $\htsm$.
1765: Theorem \ref{htm} and Lemma \ref{enough}
1766: tell us that we have found all of the relations.
1767: The proofs of the analogous statements for $\hsm$ are identical.
1768: \end{proofhtsm}
1769:
1770: How sensitive are the invariants $\htsm$ and $\hsm$?
1771: We can recover $\htm$ and $\hma$
1772: by setting $x$ to zero, hence they are at least as fine
1773: as the ordinary or $\Td$-equivariant cohomology rings.
1774: The ring $\htsm$ does {\em not} depend on coorientations,
1775: for if $\M'$ is related to $\M$ by flipping the coorientation
1776: of the $l^{\text{th}}$ hyperplane $H_k$, then the map taking
1777: % $\bd_i$ to $\bd_i$ for $i\neq l$ and
1778: $\bd_l$ to $x-\bd_l$ is an isomorphism
1779: between $\htsm$ and $H^*_{\Td\times S^1}(M')$.\footnote{The
1780: oriented matroid of a collection of nonzero vectors
1781: in a real vector space does not change when one of the vectors is negated, hence
1782: the independence of $\htsm$ on coorientations can be deduced from Remark \ref{om}.}
1783: The ring {\em does}, however, depend on $\a$, as we see in Example \ref{ts}.
1784:
1785: \begin{example}\label{ts}
1786: We compute the equivariant cohomology ring $\htsm$
1787: for the hypertoric varieties $\M_a$, $\M_b$, and $\M_c$
1788: defined by the arrangements in Figure~\ref{triplet}(a), (b), and (c),
1789: respectively.
1790: Note that each of these arrangements is smooth, hence Theorem \ref{htsm}
1791: holds over the integers, as in Remark \ref{qorz}.
1792: $$H^*_{\Td\times S^1}(\M_a)=\Z[\bd_1,\ldots,\bd_4,x]\big/
1793: \< \bd_2\bd_3, \bd_1(x-\bd_2)\bd_4, \bd_1\bd_3\bd_4 \>,$$
1794: $$H^*_{\Td\times S^1}(\M_b)=\Z[\bd_1,\ldots,\bd_4,x]\big/
1795: \< (x-\bd_2)\bd_3, \bd_1\bd_2\bd_4, \bd_1\bd_3\bd_4 \>,$$
1796: $$H^*_{\Td\times S^1}(\M_c)=\Z[\bd_1,\ldots,\bd_4,x]\big/
1797: \< \bd_2\bd_3, (x-\bd_1)\bd_2(x-\bd_4), \bd_1\bd_3\bd_4 \>.$$
1798: As we have already observed, the rings
1799: $H^*_{\Td\times S^1}(\M_a)$ and $H^*_{\Td\times S^1}(\M_b)$
1800: are isomorphic by interchanging $\bd_2$ with $x-\bd_2$.
1801: % We now show that $H^*_{\Td\times S^1}(\M_a)$ is not isomorphic
1802: % to $H^*_{\Td\times S^1}(\M_c)$.
1803: One can check that the annihilator of $\bd_2$ in $H^*_{\Td\times S^1}(\M_a)$
1804: is the principal ideal generated by $\bd_3$, while the ring $H^*_{\Td\times S^1}(\M_c)$
1805: has no degree $2$ element whose annihilator is generated by a single element of degree $2$.
1806: Hence $H^*_{\Td\times S^1}(\M_c)$ is not isomorphic to the other two rings.
1807: % Since $(x-\bd_2)(x-\bd_3)$ is the unique
1808: % degree $4$ relation in both rings, any isomorphism
1809: % between the two rings
1810: % would have to have to take $x-\bd_2$ and $x-\bd_3$
1811: % to scalar multiples of $x-\bd_2$ and $x-\bd_3$, possibly in the reverse order.
1812: % One can check that the only degree $4$ classes
1813: % in $H^*_{\Td\times S^1}(\M_a)$ that annihilate $x-\bd_2$ are multiples of $x-\bd_3$,
1814: % whereas in $H^*_{\Td\times S^1}(\M_c)$, $x-\bd_2$ is killed by $\bd_1\bd_4$
1815: % and $x-\bd_3$ is killed by $(x-\bd_1)(x-\bd_4)$.
1816: % Hence $x-\bd_2$ can be taken neither to a multiple of $x-\bd_2$ nor of $x-\bd_3$,
1817: % therefore the rings cannot be isomorphic.
1818: \end{example}
1819:
1820: The ring $\hsm$, on the other hand, is sensitive to coorientations
1821: as well as the parameter $\a$, as we see in Example \ref{s}.
1822:
1823: \begin{example}\label{s}
1824: We now compute the ring $\hsm$ for
1825: $\M_a$, $\M_b$, and $\M_c$ of Figure 2.
1826: Theorem \ref{hsm} tells us that we need only to quotient the ring
1827: $\htsm$ by $\ker(\iota^*)$.
1828: For $\M_a$, the kernel of $\iota_a^*$ is generated by
1829: $\bd_1+\bd_2-\bd_3$ and $\bd_1-\bd_4$, hence we have
1830: \begin{eqnarray*}
1831: H^*_{S^1}(\M_a)&=&\Z[\bd_2,\bd_3,x]\big/
1832: \< \bd_2\bd_3, (\bd_3-\bd_2)^2(x-\bd_2), (\bd_3-\bd_2)^2\bd_3 \>\\
1833: &\cong & \Z[\bd_2,\bd_3,x]\big/\< \bd_2\bd_3, (\bd_3-\bd_2)^2(x-\bd_2), \bd_3^3 \>.\\
1834: \end{eqnarray*}
1835: Since the hyperplanes of 2(c) have the same coorientations as those
1836: of 2(a), we have $\ker\iota_b^*=\ker\iota_a^*$, hence
1837: \begin{eqnarray*}
1838: H^*_{S^1}(\M_c) &=& \Z[\bd_2,\bd_3,x]\big/
1839: \< \bd_2\bd_3, (x-\bd_3+\bd_2)^2\bd_2, (\bd_3-\bd_2)^2\bd_3 \>\\
1840: &\cong & \Z[\bd_2,\bd_3,x]\big/\< \bd_2\bd_3, (x-\bd_3+\bd_2)^2\bd_2, \bd_3^3 \>.\\
1841: \end{eqnarray*}
1842: Finally, since Figure 2(b) is obtained from 2(a) by flipping the coorientation
1843: of $H_2$, we find that $\ker(\iota_b^*)$ is generated by $\bd_1-\bd_2-\bd_3$
1844: and $\bd_1-\bd_4$, therefore
1845: $$H^*_{S^1}(\M_b)=\Z[\bd_2,\bd_3,x]\big/
1846: \< (x-\bd_2)\bd_3, (\bd_2+\bd_3)^2\bd_2, (\bd_2+\bd_3)^2\bd_3 \>.$$
1847: As in Example \ref{ts}, $H^*_{S^1}(\M_a)$ and $H^*_{S^1}(\M_c)$
1848: can be distinguished by the fact that the annihilator
1849: of $\bd_2\in H^*_{S^1}(\M_a)$ is generated by a single element of degree $2$,
1850: and no element of $H^*_{S^1}(\M_c)$ has this property.
1851: On the other hand, $H^*_{S^1}(\M_b)$ is distinguished from
1852: $H^*_{S^1}(\M_a)$ and $H^*_{S^1}(\M_c)$ by the fact that neither $x-\bd_2$ nor $\bd_3$ cubes to zero.
1853: % Each of three rings have two linear elements, unique up to scale, that multiply to zero.
1854: % In $H^*_{S^1}(\M_c)$, both of them cube to zero, while
1855: % in $H^*_{S^1}(\M_a)$ and $H^*_{S^1}(\M_b)$, only $(x-\bd_3)$ cubes to zero.
1856: % By changing coordinates, we can write
1857: % $$H^*_{S^1}(\M_a) = \Z[x,y,z]\big/\<yz, (z-y+x)^2(x-y), z^3\>$$ and
1858: % $$H^*_{S^1}(\M_b) = \Z[x,y,z]\big/\<yz, (z-y)^2(x-y), z^3\>.$$
1859: % Any isomorphism from $H^*_{S^1}(\M_a)$ to $H^*_{S^1}(\M_b)$ would have to take
1860: % $y$ and $z$ to multiples of themselves, and $x$ to an expression of the form
1861: % $\a x+\b y+ \gamma z$, with $\a\neq 0$. Then a contradiction arises from the
1862: % fact that the first degree six relation in $H^*_{S^1}(\M_a)$ is cubic in $x$,
1863: % while $H^*_{S^1}(\M_b)$ has no such nontrivial relations.
1864: % Hence all three rings are nonisomorphic.
1865: \end{example}
1866:
1867: % \begin{remark}\label{lawrence}
1868: % Theorems \ref{htsm} and \ref{hsm} can be interpreted
1869: % in light of the recent work of Hausel and Sturmfels \cite{HS}
1870: % on Lawrence toric varieties.
1871: % The Lawrence toric variety $N$ associated to the arrangement $\A$
1872: % is the K\"ahler reduction $\cot\mod\Tk$,
1873: % so that $M$ sits inside of $N$ as the complete
1874: % intersection cut out by the equation $\bar\mc(z,w)=0$.
1875: % The residual torus acting on $N$ has dimension $d+n$,
1876: % and includes the $(d+1)$-dimensional torus $\Td\times S^1$
1877: % acting on $M$, and
1878: % the inclusion of $M$ into
1879: % $N$ induces an isomorphism on $\Td\times S^1$-equivariant cohomology.
1880: % One can use geometric arguments similar to those that were applied to prove
1881: % Theorem \ref{htsm},
1882: % or the purely combinatorial approach of \cite{HS},
1883: % to show that
1884: % $$H_{T^{d+n}}^*(N)=\Q[\bd_1,\ldots,\bd_n,v_1,\ldots,v_n]\bigg/
1885: % \< \prod_{i\in S_1}\bd_i\times
1886: % \prod_{j\in S_2}v_j\hs\bigg{\vert}\hs\bigcap_{i\in S} H_i = \emptyset\>.$$
1887: % From here we can recover $\htsm = H_{\Td\times S^1}^*(N)$
1888: % by setting $\bd_i+v_i = \bd_j+v_j$ for all $i,j\leq n$.
1889: % Note that Hausel and Sturmfels' work applies to the general orbifold case.
1890: % \end{remark}
1891: \end{section}
1892:
1893: \begin{section}{The equivariant Orlik-Solomon algebra}\label{ossection}
1894: In this section we restrict our attention to smooth arrangements.
1895: When $\A$ is smooth, all of the computations of Section \ref{htcoh}
1896: hold over the integers (see Remark \ref{qorz}). Since the rings in question
1897: are torsion-free, the presentations are also valid when the coefficients
1898: are taken in the field field $\Zt$.
1899:
1900: Let $\M_{\R} \subs \M$ be the {\em real locus}
1901: $\{[z,w]\in \M\mid z,w \text{ real}\}$ of $\M$
1902: with respect to the complex structure $J_1$.
1903: The full group $\Td\times S^1$ does not act on $\M_{\R} $,
1904: but the subgroup $\Tdr\times\Zt$ does act,
1905: where $\Tdr :=\Z_2^d\subs \Td$ is the fixed point set
1906: of the involution of $\Td$ given by complex
1907: conjugation.\footnote{It is interesting to note that
1908: the real locus with respect to the complex structure $J_1$
1909: is in fact a complex submanifold with respect to $J_3$,
1910: on which $\Tdr$ acts holomorphically and $\Zt$ acts anti-holomorphically.}
1911: % one of the other complex structures on $\M$.
1912: % The action of $\Tdr$ is holomorphic because $\Tdr$ is a subgroup of $\Td$,
1913: % which preserves all of the complex structures on $\M$.
1914: % The action of $\Zt$, on the other hand, is anti-holomorphic,
1915: % i.e. it can be thought of as complex conjugation.}
1916: In this section we will study the geometry of the real locus,
1917: focusing in particular on the properties of the $\Zt$ action.
1918: The following theorem is an extension of the results of \cite{BGH} and \cite{Sc}
1919: to the noncompact case of hypertoric varieties.
1920:
1921: \begin{theorem}\label{tara}{\em\cite{HH}}
1922: Let $G = \Tdso$ or $\Td$,
1923: and $G_{\R} = \Tdr\times\Zt$ or $\Tdr$, respectively.
1924: Then we have $H^{*}_{G}(\M;\Zt)\cong
1925: H^{*}_{G_{\R}}(\M_{\R} ;\Zt)$
1926: by an isomorphism that halves the grading.\footnote{In particular,
1927: $H^{*}_{G}(\M;\Zt)$ is concentrated in even degree.}
1928: Furthermore, this isomorphism takes the cohomology classes represented
1929: by the $G$-stable submanifolds $Z_i$ and $W_i$ to those represented by
1930: the $G_{\R}$-stable submanifolds $Z_i\cap \M_{\R}$ and $W_i\cap \M_{\R}$.
1931: \end{theorem}
1932:
1933: % \begin{sketch}
1934: % Consider the injection $H^*_G(\M;\Zt)\hookto H^*_G(\M^G;\Zt)$
1935: % given by the inclusion of the fixed point set into $\M$.
1936: % % Let \(\M^{(1)}\) be the {\em one-skeleton} of $\M$, i.e. the set of
1937: % % points in $\M$ whose $G$-orbits are 0- or 1-dimensional.
1938: % % Let \(\M^{(0)}\) denote the fixed points. Let
1939: % % \(j: \M^{(1)} \hookrightarrow \M\) and \(i: \M^{(0)} \hookrightarrow \M\)
1940: % % denote the natural inclusions.
1941: % The essential idea is to show that
1942: % % the image \(j^{*}(H^*_{G}(\M^{(1)})
1943: % % \subset H^{*}_{G}(\M^{(0)})\) is the same as the image
1944: % a class in $H^*_G(\M^G;\Zt)$ extends over $\M$ if and only
1945: % if it extends to the set of points on which $G$ acts with
1946: % a stabilizer of codimension at most $1$,
1947: % and then to show that a similar
1948: % statement in $G_{\R}$-equivariant cohomology
1949: % also holds for the real locus $\M_{\R} $ with its $G_{\R}$
1950: % action. One then uses a canonical isomorphism $H^2_G(pt,\Z_2) \cong
1951: % H^1_{G_{\R}}(pt,\Z_2)$ to give the result.
1952: %
1953: % The key to the proof is a noncompact $G_{\R}$ version
1954: % of the proposition in
1955: % \cite{TW} stating that the $G_{\R}$-equivariant Euler class of
1956: % the negative normal bundle of a critical point $p$
1957: % is not a zero divisor, which can be shown explicitly
1958: % using a local normal form for the actions of $G$ and $G_{\R}$.
1959: % % Using a local normal form theorem for
1960: % % Hamiltonian $G$-manifolds equipped with a compatible
1961: % % antisymplectic involution $\sigma$ and properties of
1962: % % the $G$ weights and $G_{\R}$ weights at the fixed points,
1963: % % one can compute
1964: % % explicitly the $G_{\R}$ equivariant Euler class and verify
1965: % %that it is not a zero divisor.
1966: % The proposition then follows
1967: % from standard $G_{\R}$ versions of the Thom isomorphism theorem
1968: % with coefficients in $\Z_2$.
1969: % Since a component of the
1970: % moment map is proper, bounded below, and has finitely many
1971: % fixed points, one can then check that the inductive argument, given
1972: % in Section 3 of \cite{TW}
1973: % to complete the proof of \cite[Thm 1]{TW} also holds in this case.
1974: % \end{sketch}
1975:
1976: Consider the restriction $\Psi$ of the hyperk\"ahler moment map
1977: $\mr\oplus\mc$ to $\M_{\R} $. Since $z$ and $w$ are real
1978: for every $[z,w]\in \M_{\R} $, the map $\mc$
1979: takes values in $\td_{\R}\subs\td_{\C}$, which we will identify
1980: with $i\R^d$, so that $\Psi$ takes values
1981: in $\R^d\oplus i\R^d\cong\C^d$.
1982: Note that $\Psi$ is $\Zt$-equivariant, with $\Zt$ acting on
1983: $\Cn$ by complex conjugation.
1984:
1985: \begin{lemma}
1986: The map $\Psi:{\M_{\R}}\to\C^d$ is surjective,
1987: and the fibers are the orbits of $\Tdr$.
1988: The stabilizer of a point $p\in {\M_{\R}}$ has order
1989: $2^r$, where $r$ is the number of hyperplanes in the complexified
1990: arrangement $\A_{\C}$ containing the point $\Psi(p)$.
1991: \end{lemma}
1992:
1993: \begin{proof}
1994: For any point $a+bi\in\C^d$, choose a point $[z,w]\in \M$ such that
1995: $\mr[z,w]=a$ and $\mc[z,w]=b$. After moving $[z,w]$ by an element of $\Td$
1996: we may assume that $z$ and $w$ are real, hence we may assume
1997: that $[z,w]\in {\M_{\R}}$.
1998: Then $$\Psi^{-1}(a+bi)=\mur^{-1}(a)\cap\muc^{-1}(b)\cap {\M_{\R}}
1999: = \Td[z,w]\cap {\M_{\R}} = \Tdr[z,w].$$
2000: The second statement follows easily from \cite[3.1]{BD}.
2001: \end{proof}
2002:
2003: Let $Y\subs {\M_{\R}}$ be the locus of points on which $\Tdr$
2004: acts freely, i.e. the preimage under $\Psi$ of the space
2005: $\MA:=\bomp$.
2006: The inclusion map $Y\hookto {\M_{\R}}$
2007: induces maps backward on cohomology, which we will denote
2008: $$\phi:\hr\to H^*_{\Tdr}(Y;\Zt)\hspace{15pt}\text{and}\hspace{15pt}
2009: \phi_2:\hrs\to H^*_{\Tdr\times\Zt}(Y;\Zt).$$
2010: By Theorem \ref{tara}, we have
2011: $$\hr\cong H^{*}_{\Td}(\M;\Zt)\hspace{15pt}\text{and}\hspace{15pt}
2012: \hrs\cong H^{*}_{\Tdso}(\M;\Zt).$$
2013: Furthermore, since $\Tdr$ acts freely on $Y$ with quotient $\MA$, we have
2014: $$H^*_{\Tdr}(Y;\Zt)\cong H^*(\MA;\Zt)\hspace{15pt}\text{and}\hspace{15pt}
2015: H^*_{\Tdr\times\Zt}(Y;\Zt)\cong H_{\Zt}^*(\MA;\Zt),$$ hence we may write
2016: $$\phi:H^{*}_{\Td}(\M;\Zt)\to H^*(\MA;\Zt)\hspace{15pt}\text{and}\hspace{15pt}
2017: \phi_2:H^{*}_{\Tdso}(\M;\Zt)\to H_{\Zt}^*(\MA;\Zt).$$
2018: The ring $H^*(\MA;\Zt)$ is a classical invariant of the arrangement $\A$ known as
2019: the {\em Orlik-Solomon algebra} (with coefficients in $\Zt$),
2020: and is denoted by $A(\A;\Zt)$ \cite{OT}.
2021: The ring $H^*_{\Zt}(\MA;\Zt)$ was introduced in \cite{HP1} and further studied
2022: in \cite{Pr}. We call it the {\em equivariant Orlik-Solomon algebra}
2023: and denote it by $A_2(\MA;\Zt)$.
2024:
2025: \begin{proposition}\label{ztf}{\em\cite[2.4]{Pr}}
2026: The space $\MA$ is $\Zt$-equivariantly formal, i.e. $A_2(\A;\Zt)$ is a free module
2027: over $\Zt[x]=\hz$.
2028: \end{proposition}
2029:
2030: \begin{theorem}\label{surj}
2031: Both $\phi$ and $\phi_2$ are surjective, with kernels
2032: $$\ker\phi = \Big\la\bd_i^2\,\bigmid\, i\in\otn\Big\ra\hspace{10pt}\text{and}
2033: \hspace{10pt}\ker\phi_2 = \Big\la\bd_i\,(x-\bd_i)\,\bigmid\, i\in\otn\Big\ra.$$
2034: \end{theorem}
2035:
2036: \begin{proof}
2037: Theorem \ref{tara} tells us that $\phi_2(\bd_i)$ is represented in
2038: $H^*_{\Tdr\times\Zt}(Y;\Zt)$ by the submanifold $Z_i\cap Y$, and likewise
2039: $\phi_2(x-\bd_i)$ by the submanifold $W_i\cap Y$. Since $\mur(Z_i\cap W_i)\subs H_i$,
2040: we have $Z_i\cap W_i\cap Y=\emptyset$, hence $\bd_i(x-\bd_i)$ lies in the kernel of
2041: $\phi_2$ (and therefore $\bd_i^2$ lies in the kernel of $\phi$).
2042:
2043: By Proposition \ref{ztf} and a pair of formal arguments
2044: identical to those of Lemmas \ref{surjective} and \ref{enough}, it is sufficient
2045: to prove Theorem \ref{surj} only for $\phi$.
2046: Quotienting $Z_i\cap Y$ by $\Tdr$, we find that $\phi(\bd_i)$ is represented in $A(\A;\Zt)$
2047: by the submanifold
2048: $$\{v\in\MA\mid v\cdot a_i + r_i \in\R^-\}.$$
2049: The standard presentation of $A(\A;\Zt)$ (see, for example, \cite{OT}) says that
2050: these classes generate the ring, and that all relations between them are generated
2051: by the monomials of Theorem \ref{hm} and $\bd_i^2$ for all $i$.
2052: \end{proof}
2053:
2054: \vspace{-\baselineskip}
2055: \begin{remark}
2056: Theorems \ref{htsm} and \ref{surj} combine to give a presentation of the equivariant
2057: Orlik-Solomon algebra $A_2(\A;\Zt)$ in the case where $\A$ is rational, simple, and smooth.
2058: This presentation first appeared in \cite{HP1}. In \cite{Pr}, we generalize this
2059: presentation to arbitrary real hyperplane arrangements, and in fact to arbitrary
2060: pointed oriented matroids.
2061: \end{remark}
2062:
2063: \begin{remark}
2064: The ring $A_2(\A;\Zt)$ is a deformation over the affine line $\operatorname{Spec}\Zt[x]$
2065: from the ordinary Orlik-Solomon algebra $A(\A;\Zt)$ to the Varchenko-Gel$'$fand ring
2066: $VG(\A;\Zt)$ of locally constant $\Zt$-valued functions on the real points of $\MA$ \cite{Pr}.
2067: While the rings $A(\A;\Zt)$ and $VG(\A;\Zt)$ depend only on the matroid associated
2068: to $\A$, the deformation $A_2(\A;\Zt)$ depends on the richer structure
2069: of the pointed oriented matroid (see Remark \ref{om}).
2070: \end{remark}
2071:
2072: \begin{example}\label{first}
2073: Consider the arrangements $\A_a$ and $\A_c$ in Figure 2(a) and 2(c).
2074: By Theorem \ref{surj} and Example \ref{ts} we have
2075: $$H^*_{\Zt}(\mathcal{M}(\A_a);\Zt)\cong
2076: \Zt [\bd_1,\ldots,\bd_4,x]\bigg/
2077: \< \begin{array}{c}
2078: \bd_1(x-\bd_1), \bd_2(x-\bd_2), \bd_3(x-\bd_3), \bd_4(x-\bd_4),\\
2079: \bd_2\bd_3, \bd_1(x-\bd_2)\bd_4, \bd_1\bd_3\bd_4
2080: \end{array}\>$$
2081: and
2082: $$H^*_{\Zt}(\mathcal{M}(\A_c);\Zt)\cong
2083: \Zt [\bd_1,\ldots,\bd_4,x]\bigg/
2084: \< \begin{array}{c}
2085: \bd_1(x-\bd_1), \bd_2(x-\bd_2), \bd_3(x-\bd_3), \bd_4(x-\bd_4),\\
2086: \bd_2\bd_3, (x-\bd_1)\bd_2(x-\bd_4), \bd_1\bd_3\bd_4
2087: \end{array}\>.$$
2088: The map $f:H^*_{\Zt}(\mathcal{M}(\A_a);\Zt)\to H^*_{\Zt}(\mathcal{M}(\A_b);\Zt)$
2089: given by
2090: $$f(\bd_1) = \bd_1+\bd_2, f(\bd_2) = \bd_2+\bd_3+x, f(\bd_3) = \bd_3,
2091: f(\bd_4) = \bd_2+\bd_4,\text{ and }f(x)=x$$
2092: is an isomorphism of graded $\Zt[x]$-algebras, despite the fact that the
2093: oriented matroids of the two arrangements differ.\footnote{We thank
2094: Graham Denham for finding this isomorphism.}
2095: \end{example}
2096:
2097: \begin{example}\label{second}
2098: Now consider the arrangements $\A_a'$ and $\A_c'$ obtained from $\A_a$ and $\A_c$
2099: by adding a vertical line on the far left, as shown in Figure \ref{vertical}.
2100: \begin{figure}[h]
2101: \begin{center}
2102: \psfrag{Ha'}{$\A_a'$}
2103: \psfrag{Hc'}{$\A_c'$}
2104: \includegraphics{htv3.eps}
2105: \caption{Adding a vertical line to $\A_a$ and $\A_c$.}\label{vertical}
2106: \end{center}
2107: \end{figure}
2108: Again by Theorem \ref{surj}, we have
2109: $$H^*_{\Zt}(\mathcal{M}(\A_a');\Zt)\cong
2110: \Zt [\bd_1,\ldots,\bd_4,x]\bigg/
2111: \< \begin{array}{c}
2112: \bd_1(x-\bd_1), \bd_2(x-\bd_2), \bd_3(x-\bd_3), \bd_4(x-\bd_4),\\ \bd_5(x-\bd_5),
2113: \bd_2\bd_3, (x-\bd_1)\bd_5, \bd_1(x-\bd_2)\bd_4,\\ \bd_1\bd_3\bd_4, (x-\bd_2)\bd_4\bd_5, \bd_3\bd_4\bd_5
2114: \end{array}\>$$
2115: and
2116: $$H^*_{\Zt}(\mathcal{M}(\A_c');\Zt)\cong
2117: \Zt [\bd_1,\ldots,\bd_4,x]\bigg/
2118: \< \begin{array}{c}
2119: \bd_1(x-\bd_1), \bd_2(x-\bd_2), \bd_3(x-\bd_3), \bd_4(x-\bd_4),\\ \bd_5(x-\bd_5),
2120: \bd_2\bd_3, (x-\bd_1)\bd_5, (x-\bd_1)\bd_2(x-\bd_4),\\ \bd_1\bd_3\bd_4, (x-\bd_2)\bd_4\bd_5, \bd_3\bd_4\bd_5
2121: \end{array}\>.$$
2122: We have used Macaulay 2 \cite{M2} to check that the annihilator of the element
2123: $\bd_2\in H^*_{\Zt}(\mathcal{M}(\A_a');\Zt)$ is generated by two linear elements
2124: (namely $\bd_3$ and $x-\bd_2$) and nothing else, while there is no element
2125: of $H^*_{\Zt}(\mathcal{M}(\A_c');\Zt)$ with this property.
2126: Hence the two rings are not isomorphic.
2127: \end{example}
2128: \end{section}
2129:
2130: \begin{section}{Cogenerators}\label{cogs}
2131: Consider the K\"ahler Kirwan map
2132: $$\ka:\Sym\tkd\cong H^*_{\Tk}(\C^n)\to H^*(\Xa_{\a})$$
2133: induced by the $\Tk$-equivariant
2134: inclusion of $\mu^{-1}(\a)$ into $\Cn$.
2135: In this section we would like to consider simultaneously the Kirwan
2136: maps $\ka$ for many different values of $\a$, so almost all of the
2137: notation that we use will have a subscript or superscript indicating
2138: the parameter $\a\in\tkd$ or a lift $r\in\tnd$.
2139: An exception to this rule will be the
2140: hyperk\"ahler Kirwan map $$\k:\Sym\tkd\to H^*(\Ma_{\a}),$$
2141: which, by Lemma \ref{placement} or Theorem \ref{hm},
2142: is independent of our choice of simple $\a\in\tkd$.
2143: The main result of this section is the following.
2144:
2145: \begin{theorem}\label{int}
2146: The kernel of the hyperk\"ahler Kirwan map $\k$ is equal to the intersection over all simple
2147: $\a$ of the kernels of the K\"ahler Kirwan maps $\ka$.
2148: \end{theorem}
2149:
2150: \begin{remark}
2151: Konno \cite[7.6]{K2} proves an analogous theorem about the kernels
2152: of the Kirwan maps to the cohomology rings of polygon and
2153: hyperpolygon spaces. We may therefore conjecture a generalization
2154: of Theorem \ref{int} in which $\Tk$ is replaced by an arbitrary compact
2155: group $G$. Note that our proof of Theorem \ref{int}
2156: depends strongly on the combinatorics associated to hypertoric varieties.
2157: \end{remark}
2158:
2159: We approach Theorem \ref{int} by describing the kernels of $\k$ and $\ka$ not
2160: in terms of generators, but rather in terms of cogenerators.
2161: Given an ideal $\mathcal{I}\subs\Sym\tkd$, a set of {\em cogenerators} for $\mathcal{I}$
2162: is a collection of polynomials $\{f_i\}\subs\Sym\tk$ such that
2163: $$\mathcal{I}=\{\bd\in\Sym\tkd\mid\bd\cdot f_i = 0\text{ for all }i\}.$$
2164:
2165: The volume function $\Vol\Dr$ is locally polynomial in $r$.
2166: More precisely, for every simple $r\in\tnd$, there exists a degree $d$ polynomial
2167: $\Pr\in\Sym^d\tn$ such that for every simple $s\in\tnd$ lying in the
2168: same connected component of the set of simple elements as $r$,
2169: we have $$\Vol\D^{s}=\Pr\!(s).$$ We will refer to $\Pr$ as the
2170: {\em volume polynomial} of $\Dr$. The fact that the volume of a polytope
2171: is translation invariant tells us that $\Pr$ lies in the image of the inclusion
2172: $\i:\Sym^d\tk\hookto\Sym^d\tn$.
2173:
2174: \begin{theorem}\label{toric}{\em\cite{GS,KP}}
2175: Let $r\in\tnd$ be a simple element with $\i^*(r)=\a$. Then
2176: % $$\ker (\ka\circ\i^*) = \Ann\{\Pr\} = \big\{\bd\in\Sym\tnd \mid \bd\cdot\Pr = 0\big\}$$ and
2177: $$\ker\ka = \Ann\{\i^{-1}\Pr\} =
2178: \big\{\bd\in\Sym\tkd \mid \bd\cdot(\i^{-1}\Pr) = 0\big\}.$$
2179: \end{theorem}
2180:
2181: A similar description of the cohomology ring of a hypertoric variety
2182: is given in \cite{HS}. For any subset
2183: $A\subs\otn$, consider the polyhedron $\DrA$ introduced in Section \ref{htcore}.
2184: If $\DrA$ is nonempty, then it is bounded if and only if the vectors
2185: $\{\eps_1(A)a_1,\ldots,\eps_n(A)a_n\}$ span $\td$ over the non-negative
2186: real numbers,
2187: where $\eps_i(A) = (-1)^{|A\cap\{i\}|}$.
2188: We call such an $A$ {\em admissible}. For all admissible $A$,
2189: there exists a degree $d$ polynomial
2190: $\PrA\in\Sym^d\tn$ such that for every simple $s\in\tnd$
2191: lying in the same connected component of the set of simple elements as $r$,
2192: we have $$\Vol\D^{s}_A=\PrA(s).$$
2193: Once again, the translation invariance of volume implies that
2194: $\PrA$ lies in the image of the inclusion
2195: $\i:\Sym^d\tk\hookto\Sym^d\tn$.
2196: Consider the linear span
2197: $$U^r = \Q\big\{\PrA \mid A\text{ admissible}\big\}.$$
2198:
2199: \begin{theorem}\label{hypertoric}{\em\cite{HS}}
2200: Let $r\in\tnd$ be a simple element with $\i^*(r)=\a$. Then
2201: % $$\ker (\hk\circ\i^*) = \Ann U^r = \big\{\bd\in\Sym\tnd \mid
2202: % \bd\cdot P = 0\text{ for all }P\in U^r\big\}$$ and
2203: $$\ker\k = \Ann \,\i^{-1}U^r = \big\{\bd\in\Sym\tkd \mid
2204: \bd\cdot \i^{-1}P = 0\text{ for all }P\in U^r\big\}.$$
2205: \end{theorem}
2206:
2207: \begin{remark}\label{wow}
2208: It is clear from the statement of Theorem \ref{hypertoric}
2209: that $H^*(\Ma)$ does not depend on the coorientations of the hyperplanes
2210: $\{H_1^r,\ldots,H_n^r\}$, as has been observed in Lemma \ref{coorientation}
2211: and throughout Section \ref{htcoh}. Indeed,
2212: the polynomials $\PrA$ for $A$ admissible
2213: are simply the volume polynomials of the maximal regions of $|\A|$.
2214: What is not clear from this presentation is the independence of $H^*(\Ma)$
2215: on $\a$. In other words, it is a nontrivial
2216: fact that the vector space $U^r$ is independent of the parameter $r\in\tnd$.
2217: \end{remark}
2218:
2219: \begin{proofint}
2220: The statement of Theorem \ref{int} equates the kernel of $\k$,
2221: which is cogenerated by the vector
2222: space $\i^{-1}U^r$, with the intersection of the kernels of the maps $\ka$,
2223: each of which is cogenerated by the element $\i^{-1}P^r$. Intersection of ideals
2224: corresponds to linear span on the level of cogenerators, hence we have
2225: $$\bigcap_{\a\text{ simple}}\ker\ka = \Ann \,\i^{-1}V,\hspace{15pt}\text{where}
2226: \hspace{15pt}V = \Q\big\{\Pr\mid r\text{ simple}\}.$$
2227: Our plan is to show that $V = U^r$ for any simple $r$.
2228: Recall that the assignment of $\Pr$ to $r$ is locally constant on the set
2229: of simple elements of $\tnd$, hence $V$ is finite-dimensional.
2230: Since $\Pr\in U^r$ and $U^r$ does not depend on $r$ (see Remark \ref{wow}),
2231: it is clear that $V\subs U^r$. Thus to prove Theorem \ref{int}, it will suffice
2232: to prove the opposite inclusion, as stated below.
2233:
2234: \begin{proposition}\label{comb}
2235: We have $\PrA\in V$ for every admissible $A\subs\otn$.
2236: \end{proposition}
2237:
2238: Let $\F$ be the infinite dimensional vector space consisting of all real-valued
2239: functions on $\tdd$, and let $\Fbd$ be the subspace consisting of functions
2240: with bounded support. For all subsets $A\subs\otn$, let
2241: $$\WA = \Q\big\{\odra\mid r\text{ simple}\big\}$$
2242: be the subspace of $\F$ consisting of finite linear combinations
2243: of characteristic functions of polyhedra $\DrA$, and let
2244: $$\WAbd = \WA\cap\Fbd.$$
2245: Note that $\WAbd=\WA$ if and only if $A$ is admissible.
2246:
2247: \begin{lemma}\label{key}
2248: For all $A,A'\subs\otn$, $\WAbd = W_{A'}^{bd}$.
2249: \end{lemma}
2250:
2251: \begin{proof}
2252: We may immediately reduce to the case where $A' = A\cup\{j\}$.
2253: Fix a simple $r\in\tnd$.
2254: Let $\rt\in\tnd$ be another simple element obtained from $r$ by putting
2255: $\rt_i=r_i$ for all $i\neq j$, and $\rt_j = N$ for some $N\gg 0$.
2256: Then $\D_A^r\subs\D_A^{\rt}$, and
2257: \begin{eqnarray*}
2258: \D_{A}^{\rt} \smallsetminus \D_{A}^{r} &=&
2259: \big\{v\in\tdd\mid \eps_i(A')(v\cdot a_i+r_i)\geq 0\text{ for all }i\leq n
2260: \text{ and } v\cdot a_j + N\geq 0\}\\
2261: &=& \D_{A'}^{r} \cap G_j^{N}.
2262: \end{eqnarray*}
2263: Suppose that $f\in\Fbd$ can be written as a linear combination
2264: of functions of the form $\mathbf{1}_{\D_{A'}^{r}}$. Choosing $N$ large enough
2265: that the support of $f$ is contained in $G_j^{N}$,
2266: the above computation shows that $f$ can be written
2267: as a linear combination of functions of the form $\odra$,
2268: hence $W_{A'}^{bd}\subs W_{A}^{bd}$. The reverse inclusion is obtained
2269: by an identical argument.
2270: \end{proof}
2271:
2272: \vspace{-\baselineskip}
2273: \begin{example}\label{pictures}
2274: Suppose that we want to write the characteristic function for the
2275: upper triangle $\Delta_{\{1,4\}}$
2276: in Figure~\ref{triplet}(c) as linear combination
2277: of characteristic functions of the shaded regions obtained by translating
2278: the hyperplanes in any possible way.
2279: Since $\{1,4\}$ has two elements, the procedure described
2280: in Lemma \ref{key} must be iterated twice, and the result will have a total
2281: of $2^2=4$ terms, as illustrated in Figure \ref{char}.
2282: The first iteration exhibits $\mathbf{1}_{\Delta_{\{1,4\}}}$ as an element
2283: of $W_{\{4\}}^{bd}$ by expressing it as the difference of the characteristic
2284: functions of two (unbounded) regions.
2285: With the second iteration, we attempt to express each of these two characteristic functions
2286: as elements of $W_{\{1,4\}}^{bd} = W_{\{1,4\}}$.
2287: This attempt must fail, because each of the two functions that we try
2288: to express has unbounded support. But the failures cancel out, and we succeed
2289: in expressing the {\em difference} as an element of $W_{\{1,4\}}$.
2290: \begin{figure}[h]
2291: \begin{center}
2292: \includegraphics[height=140mm]{newchar.eps}
2293: \caption{An equation of characteristic functions.
2294: We write two numbers next to each hyperplane: the first is the
2295: index $i\in\{1,\ldots,4\}$, and the second is the parameter $r_i$
2296: specifying the distance from the origin (denoted by a black dot) to $H_i$.
2297: The two iterations of Lemma \ref{key} have produced two undetermined
2298: large numbers, which we call $N$ and $N'$.}\label{char}
2299: \end{center}\end{figure}
2300: \end{example}
2301:
2302: \begin{proofcomb}
2303: By Lemma \ref{key}, we may write
2304: $$\odra = \sum_{j=1}^m\eta_j \mathbf{1}_{\D^{r(j)}}$$
2305: for any simple $r$ and admissible $A$, where $\eta_j\in\Z$ and
2306: $r(j)$ is a simple element of $\tnd$ for all $j\leq m$.
2307: Taking volumes of both sides of the equation, we have
2308: \begin{equation}\label{polys}
2309: \PrA(r) = \sum_{j=1}^m\eta_j P^{r(j)}\big(r(j)\big).
2310: \end{equation}
2311: Furthermore, we observe from the proof of Lemma \ref{key}
2312: that for all $j\leq m$ and all $i\leq n$, the $i^{\text{th}}$
2313: coordinate $r_i(j)$ of $r(j)$ is either equal to $r_i$, or to
2314: some large number number $N\gg 0$. The Equation~\eqref{polys}
2315: still holds if we wiggle these large numbers a little bit,
2316: hence the polynomial $P^{r(j)}$ must be independent of the variable
2317: $r_i(j)$ whenever $r_i(j)\neq r_i$. Thus we may substitute
2318: $r$ for each $r(j)$ on the right-hand side, and we obtain the equation
2319: \begin{equation*}
2320: \PrA(r) = \sum_{j=1}^m\eta_j P^{r(j)}(r).
2321: \end{equation*}
2322: This equation clearly holds in a neighborhood of $r$,
2323: hence we obtain an equation of polynomials
2324: $$\PrA = \sum_{j=1}^m\eta_j P^{r(j)}.$$
2325: This completes the proof of
2326: Proposition \ref{comb}, and therefore also of Theorem \ref{int}.
2327: \end{proofcomb}
2328: \end{proofint}
2329:
2330: \vspace{-\baselineskip}
2331: \begin{example}
2332: Let's see what happens when we take volume polynomials in the equation
2333: of Figure \ref{char}. The two polytopes on the top line have different
2334: volumes, but the same volume polynomial, hence these two terms cancel.
2335: We are left with the equation
2336: $$P_{\{1,4\}}^{(0,1,1,0)} = P^{(0,1,1,0)} - P^{(N,1,1,0)},$$
2337: which translates as
2338: $$\half\(-x_1+x_2-x_4\)^2 = \half\(x_1+x_3+x_4\)^2 -
2339: \(x_2+x_3\)\(x_1+x_4+\half x_3 - \half x_2\).$$
2340: \end{example}
2341: \end{section}
2342: \end{chapter}
2343:
2344: \begin{chapter}{Abelianization}\label{abelianization}
2345: Let $X$ be a symplectic manifold equipped with a hamiltonian
2346: action of a compact Lie group $G$. Let $T\subs G$
2347: be a maximal torus, let $\Delta\subset\tdu$
2348: be the set of roots\footnote{Not to be confused with the polyhedron $\Delta$
2349: of Chapter \ref{ht}.} of $G$,
2350: and let $W=N(T)/T$ be the Weyl group of $G$.
2351: If $\mu:X\to\gd$ is a moment map for the action of $G$, then $pr\circ\mu:X\to\tdu$
2352: is a moment map for the action of $T$, where $pr:\gd\to\tdu$ is the standard projection.
2353: Suppose that $0\in\gd$ and $0\in\tdu$ are regular values for the two moment maps.
2354: If the symplectic quotients
2355: $$X\mod G = \mu^{-1}(0)/G\hspace{15pt}\text{and}
2356: \hspace{15pt}X\mod T=(pr\circ\mu)^{-1}(0)/T$$
2357: are both compact, then Martin's theorem \cite[Theorem A]{Ma}
2358: relates the cohomology of $X\mod G$ to the cohomology of $X\mod T$.
2359: Specifically, it says that $$H^*(X\mod G)\cong\frac{H^*(X\mod T)^W}{ann(e_0)},$$
2360: where $$e_0 = \prod_{\a\in\Delta}\a\in\left(\operatorname{Sym}\tdu\right)^W
2361: \cong H^*_T(pt)^W,$$ which acts naturally on
2362: $H^*(X\mod T)^W \cong H^*_T(\mu_T^{-1}(0))^W$.
2363: In the case where $X$ is a complex vector space and $G$ acts linearly
2364: on $X$, a similar result was obtained by Ellingsrud and Str\o mme \cite{ES}
2365: using different techniques.
2366:
2367: Our goal is to state and prove an analogue of this theorem for
2368: hyperk\"ahler quotients. There are two main obstacles to this goal.
2369: First, hyperk\"ahler quotients are rarely compact.
2370: The assumption of compactness in Martin's theorem is crucial
2371: because his proof involves integration.
2372: Our answer to this problem is to work
2373: with equivariant cohomology
2374: of {\em circle compact} manifolds, by which we mean oriented manifolds
2375: with an action of $\so$ such that the fixed point set is oriented and compact.
2376: Using the localization theorem of Atiyah-Bott \cite{AB} and Berline-Vergne \cite{BV},
2377: as motivation, we show that integration in rationalized $\so$-equivariant cohomology
2378: of circle compact manifolds can be defined in terms of integration
2379: on their fixed point sets. Section~\ref{thepush} is devoted to
2380: making this statement precise by defining a well-behaved
2381: push forward in the rationalized $\so$-equivariant cohomology
2382: of circle compact manifolds.
2383:
2384: The second obstacle is that Martin's result uses surjectivity
2385: of the K\"ahler Kirwan map from $H^*_G(X)$ to $H^*(X\mod G)$ \cite{Ki}.
2386: The analogous map for circle compact
2387: hyperk\"ahler quotients is conjecturally surjective, but
2388: only a few special cases are known (see Theorems \ref{hm} and \ref{hpsurj}, and
2389: Remarks \ref{generation} and \ref{horn}).
2390: Our approach is to assume that the rationalized Kirwan map is
2391: surjective, which is equivalent to saying that the cokernel
2392: of the non-rationalized Kirwan map
2393: $$\K_G:H^*_{\so\times G}(M)\to\hmg$$
2394: is torsion as a module over $\hp$. This is a weaker assumption
2395: than surjectivity of $\K_G$; in particular, we show in Section
2396: \ref{quiver} that this assumption holds for quiver varieties,
2397: as a consequence of the work of Nakajima.
2398:
2399: Under this assumption,
2400: Theorem~\ref{main} computes the rationalized equivariant cohomology
2401: of $M\mmod G$ in terms of that of $M\mmod T$.
2402: We show that, at regular values of the hyperk\"ahler moment maps,
2403: $$\hhso(M\mmod G)\cong\frac{\hhso(M\mmod T)^W}{ann(e)},$$
2404: where $\hhso$ denotes rationalized equivariant cohomology
2405: (see Definition \ref{rat}), and
2406: $$e = \Pd_{\a\in\Delta}\a(x-\a) \in (\operatorname{Sym}\tdu)^W\otimes\Q[x]
2407: \cong H^*_{\so\times T}(pt)^W\subs\widehat{H}^*_{\so\times T}(pt)^W.$$
2408: Theorem~\ref{ordinary}
2409: describes the image of the non-rationalized Kirwan map
2410: in a similar way:
2411: $$\hso(M\mmod G)\supseteq\Im(\K_G)\cong\frac{\imkt^W}{ann(e)},$$
2412: where $\K_T:H^*_{\so\times T}(M)\to \hso(M\mmod T)$ is the Kirwan
2413: map for the abelian quotient.
2414: In many situations, such as when $M=\cot$,
2415: $\K_T$ is known to be surjective (Theorem \ref{hm}).
2416:
2417: This Chapter is a reproduction of \cite[\S 1-3]{HP}.
2418:
2419: \begin{section}{Integration}\label{thepush}
2420: The localization theorem of Atiyah-Bott \cite{AB} and Berline-Vergne \cite{BV} says that
2421: given a manifold $M$ with a circle action, the restriction map from
2422: the circle-equivariant cohomology of $M$ to the circle-equivariant
2423: cohomology of the fixed point set $F$ is an isomorphism modulo torsion.
2424: In particular, integrals on $M$ can be computed in terms of integrals on $F$.
2425: If $F$ is compact, it is possible to use the Atiyah-Bott-Berline-Vergne formula
2426: to {\em define} integrals on $M$.
2427:
2428: We will work in the category of {\em circle compact}
2429: manifolds, by which we mean oriented $\so$-manifolds with compact
2430: and oriented fixed point sets.
2431: Maps between circle compact
2432: manifolds are required to be equivariant.
2433: % Our goal is to define a cohomological pushforward in the category
2434: % of circle compact manifolds, and derive some of its basic properties.
2435:
2436: \begin{definition}\label{rat}
2437: Let $\hhp = \Q(x)$, the rational function field of $\hp\cong \Q[x]$.
2438: For a circle compact manifold $M$, let $\hhm = \hm\otimes\hhp$,
2439: where the tensor product is taken over the ring $\hp$. We call $\hhm$
2440: the {\em rationalized} $\so$-equivariant cohomology of $M$.
2441: Note that because $\operatorname{deg}(x)=2$, $\hhm$ is supergraded,
2442: and supercommutative with respect to this supergrading.
2443: \end{definition}
2444:
2445: An immediate consequence of \cite{AB} is that restriction gives an isomorphism
2446: \begin{equation}\label{isomorphic}\hhm\cong \hhso(F)\cong H^*(F)\otimes_{\Q}\hhp,\end{equation}
2447: where $F=M^{S^1}$ denotes
2448: the compact fixed point set of $M$. In particular
2449: $\hhm$ is a finite dimensional vector space over $\hhp$,
2450: and trivial if and only if $F$ is empty.
2451:
2452: Let $i:N\hookto M$ be a closed embedding.
2453: There is a standard notion of proper pushforward
2454: $$i_*:\hso(N)\to\hm$$ given by the formula $i_*=r\circ\Phi$,
2455: where $r:\hso(M,M\setminus N)\to\hso(M)$ is the restriction map,
2456: and $\Phi:\hso(N)\to\hso(M,M\setminus N)$ is the Thom isomorphism.
2457: We will also denote the induced map $\hhso(N)\to\hhm$ by $i_*$.
2458: Geometrically, $i_*$ can be understood as the inclusion
2459: of cycles in Borel-Moore homology.
2460:
2461: This map satisfies two important formal properties
2462: \cite{AB}:
2463: \begin{equation}\label{functor}
2464: \text{Functoriality: } (i\circ j)_* = i_*\circ j_*
2465: \end{equation}
2466: \begin{equation}\label{module}
2467: \text{Module homomorphism: } i_*(\ga\cdot i^*\a)
2468: =i_\ga\cdot\a
2469: \text{ for all }\a\in\hhm, \ga\in\hhn.
2470: \end{equation}
2471: We will denote the Euler
2472: class $i^*i_*(1)\in\hhso(N)$ by $e(N)$.
2473: If a class $\ga\in\hhn$ is in the image of $i^*$,
2474: then property (\ref{module}) tells us that
2475: $i^*i_*\ga = e(N)\ga$. Since the pushforward construction is
2476: local in a neighborhood
2477: of $N$ in $M$, we may assume that $i^*$ is surjective, hence this identity
2478: holds for all $\ga\in\hhn$.
2479:
2480: Let $F= M^{\so}$ be the fixed point set of $M$.
2481: Since $M$ and $F$ are each oriented, so is the normal bundle
2482: to $F$ inside of $M$. The following result is
2483: standard, see e.g. \cite{Ki}.
2484:
2485: \begin{lemma}
2486: The Euler class $e(F)\in\hhso(F)$ of the normal bundle to $F$ in $M$
2487: is invertible.
2488: \end{lemma}
2489:
2490: \begin{proof}
2491: Let $\{F_1,\ldots,F_d\}$ be the connected components of $F$.
2492: Since $\hhso(F)\cong\bigoplus\hhso(F_i)$ and $e(F)=\oplus e(F_i)$,
2493: our statement is equivalent to showing that $e(F_i)$ is invertible for all $i$.
2494: Since $\so$ acts trivially on $F_i$, $\hhso(F_i)\cong H^*(F_i)\otimes_{\Q}\hhp$.
2495: We have $e(F_i) = 1\otimes ax^k + nil$, where
2496: $k= \operatorname{codim}(F_i)$, $a$ is the product of the weights of the $S^1$
2497: action on any fiber of the normal bundle, and $nil$ consists of terms
2498: of positive degree in $H^*(F_i)$. Since $F_i$ is a component of the fixed
2499: point set, $\so$ acts freely on the complement of the zero section
2500: of the normal bundle, therefore $a\neq 0$. Since $ax^k$ is invertible and
2501: $nil$ is nilpotent, we are done.
2502: \end{proof}
2503:
2504: \vspace{-\baselineskip}
2505: \begin{definition}
2506: For $\a\in\hhm$, let $$\IM\a=\IF\frac{\a|_F}{e(F)}\in\hhp.$$
2507: \end{definition}
2508:
2509: Note that this definition does not depend on our choice
2510: of orientation of $F$. Indeed, reversing the orientation of $F$
2511: has the effect of negating $e(F)$, {\em and} introducing a second factor
2512: of $-1$ coming from the change in fundamental class. These two effects
2513: cancel.
2514:
2515: For this definition to be satisfactory, we must be able to prove
2516: the following lemma, which is standard in the setting of
2517: ordinary cohomology of compact manifolds.
2518:
2519: \begin{lemma}\label{Npush}
2520: Let $i:N\hookto M$ be a closed immersion.
2521: Then for any $\a\in\hhm,\ga\in\hhso(N)$, we have
2522: $\IM\a\cdot i_*\ga = \IN i^*\a\cdot\ga$.
2523: \end{lemma}
2524:
2525: \begin{proof}
2526: Let $G=N^{\so}$, let $j:G\to F$ denote the restriction of $i$ to $G$,
2527: and let $\phi:F\to M$ and $\psi:G\to N$ denote the inclusions of $F$ and $G$
2528: into $M$ and $N$, respectively.
2529: $$\begin{CD}
2530: N @>i>> M\\
2531: @A\psi AA @AA\phi A\\
2532: G @>j>> F
2533: \end{CD}$$
2534: Then $$\IM\a\cdot i_*\ga = \IF\frac{\phi^*\a\cdot\phi^*i_*\ga}{e(F)},$$
2535: and $$\IN i^*\a\cdot\ga = \IG\frac{\psi^*i^*\a\cdot\psi^*\ga}{e(G)}
2536: =\IG\frac{j^*\phi^*\a\cdot\psi^*\ga}{e(G)}
2537: =\IF\phi^*\a\cdot j_*\left(\frac{\psi^*\ga}{e(G)}\right),$$
2538: where the last equality is simply the integration formula applied
2539: to the map $j:G\to F$ of compact manifolds \cite{AB}.
2540: Hence it will be sufficient to prove that
2541: $$\phi^*i_*\ga=e(F)\cdot j_*\left(\frac{\psi^*\ga}{e(G)}\right)\in\hhso(F).$$
2542: To do this, we will
2543: show that the difference of the two classes lies in
2544: the kernel of $\phi_*$, which we know is trivial because the composition
2545: $\phi^*\phi_*$ is given by multiplication by the invertible
2546: class $e(F)\in\hhso(F)$.
2547: On the left hand side we get
2548: $$\phi_*\phi^*i_*\ga = \phi_*(1)\cdot i_*\ga\hspace{15pt}
2549: \text{by }(\ref{module}),
2550: $$
2551: and on the right hand side we get
2552: \begin{eqnarray*}
2553: \phi_*\left(e(F)\cdot j_*\left(\frac{\psi^*\ga}{e(G)}\right)\right)
2554: &=& \phi_*\left(\phi^*\phi_*(1)
2555: \cdot j_*\left(\frac{\psi^*\ga}{e(G)}\right)\right)\\
2556: &=& \phi_*(1)\cdot \phi_*j_*\left(\frac{\psi^*\ga}{e(G)}\right)
2557: \hspace{1in}\text{ by }(\ref{module})\\
2558: &=& \phi_*(1)\cdot i_*\psi_*\left(\frac{\psi^*\ga}{e(G)}\right)
2559: \hspace{1in}\text{ by }(\ref{functor}).
2560: \end{eqnarray*}
2561: It thus remains only to show that
2562: $\ga = \psi_*\left(\frac{\psi^*\ga}{e(G)}\right)$.
2563: This is seen by applying $\psi^*$ to both sides,
2564: which is an isomorphism (working over the field $\hhp$) by \cite{AB}.
2565: % But this is simply the localization theorem of \cite{AB}, proven by applying
2566: % $\psi^*$ to both sides. (Atiyah and Bott prove that
2567: % $\psi^*:\hso(N)\to\hso(G)$ is injective
2568: % modulo torsion; by tensoring with $\hhp$, we are working modulo torsion.)
2569: \end{proof}
2570:
2571: For $\a_1,\a_2\in\hhm$, consider the
2572: symmetric, bilinear, $\hhp$-valued pairing
2573: $$\la\a_1,\a_2\ra_M=\IM\a_1\a_2.$$
2574:
2575: \begin{lemma}[Poincar\'e Duality]\label{perfect}
2576: This pairing is nondegenerate.
2577: \end{lemma}
2578:
2579: \begin{proof}
2580: Suppose that $\a\in\hhm$ is nonzero, and therefore $\phi^*\a\neq 0$.
2581: Since $F$ is
2582: compact, there must exist a class $\ga\in\hhso(F)$ such
2583: that $0\neq\IF\phi^*\a\cdot\ga=\IM\a\cdot \phi_*\ga=\la\a,\phi_*\ga\ra_M$.
2584: \end{proof}
2585:
2586: \vspace{-\baselineskip}
2587: \begin{definition}\label{push}
2588: For an arbitrary equivariant map $f:N\to M$, we may now define
2589: the pushforward $$f_*:\hhn\to\hhm$$ to
2590: be the adjoint of $f^*$ with respect to the pairings
2591: $\la\cdot,\cdot\ra_N$ and $\la\cdot,\cdot\ra_M$.
2592: This is well defined because, according to (\ref{isomorphic}),
2593: $\hhm$ and $\hhn$ are finite dimensional
2594: vector spaces over the field $\hhp$. Lemma~\ref{Npush} tells us that this definition generalizes
2595: the definition for closed immersions. Furthermore,
2596: properties (\ref{functor}) and (\ref{module}) for pushforwards along
2597: arbitrary maps
2598: are immediate corollaries of the definition.
2599: If $f$ is a projection, then $f_*$ will be given by
2600: integration along the fibers.
2601: Using the fact that
2602: every map factors through its graph as a closed immersion
2603: and a projection, we always have a geometric interpretation of the pushforward.
2604: \end{definition}
2605:
2606: As an application, let us consider the manifold $M\times M$,
2607: along with the two projections $\pi_1$ and $\pi_2$,
2608: and the diagonal map $\Diag:M\to M\times M$.
2609: Suppose that we can write $$\Diag_*(1)=\sum\pi_1^*a_i\cdot\pi_2^*b_i$$
2610: for a finite collection of classes $a_i,b_i\in\hhm$.
2611: The following Proposition will be used in Section~\ref{quiver}.
2612:
2613: \begin{proposition}\label{basis}
2614: The set $\{b_i\}$ is an additive spanning set for $\hhm$.
2615: \end{proposition}
2616:
2617: \begin{proof}
2618: For any $\a\in\hhm$, we have
2619: \begin{eqnarray*}
2620: \a &=& \id_*\id^*\a\\
2621: &=& (\pi_2\circ\Diag)_*(\pi_1\circ\Diag)^*\a\\
2622: &=& \pi_{2*}\big(\Diag_*\left(1\cdot\Diag^*\pi_1^*\a\right)\big)\\
2623: &=& \pi_{2*}\big(\pi_1^*\a\cdot\Diag_*(1)\big)\\
2624: &=& \pi_{2*}\left(\sum\pi_1^*(a_i\a)\cdot\pi_2^*b_i\right)\\
2625: &=& \sum\pi_{2*}\pi_1^*(a_i\a) \cdot b_i\\
2626: &=& \sum\la a_i,\a\ra\cdot b_i,
2627: \end{eqnarray*}
2628: hence $\a$ is in the span of $\{b_i\}$.
2629: \end{proof}
2630: \end{section}
2631:
2632: \begin{section}{Hyperk\"ahler abelianization}
2633: Let $M$ be a hyperk\"ahler manifold with a circle action, and
2634: suppose that a compact Lie group $G$ acts hyperhamiltonianly on $M$
2635: with hyperk\"ahler moment map
2636: $$\muh=\mur\oplus\muc:M\to\gd\oplus\gd_{\C},$$
2637: where $\muc$ is holomorphic with respect to the distinguished
2638: complex structure $I$.
2639: We require that the action of $G$ commute with the action of $\so$,
2640: that $\mur$ is $\so$-invariant, and that $\muc$ is $\so$-equivariant
2641: with respect to the action of $\so$ on $\gd_{\C}$ by complex multiplication.
2642:
2643: Let $T\subs G$ be a maximal torus, and let $pr:\gd\to\tdu$
2644: be the natural projection. Then $T$ acts on $M$ with hyperk\"ahler
2645: moment map
2646: $$\mubh=pr\!\circ\!\mur\,\oplus\, pr_{\C}\!\circ\!\muc:
2647: M\to\tdu\oplus\tdu_{\C}.$$
2648: Let $\xi\in\gd$ be a central element such that $(\xi,0)$ is a regular
2649: value of $\muh$ and $(pr(\xi),0)$ is a regular value of $\mubh$.
2650: Assume further that $G$ acts freely on $\muh^{-1}(\xi,0)$,
2651: and $T$ acts freely on $\mubh^{-1}(pr(\xi),0)$.\footnote{We make this
2652: simplifying assumption in order to talk about manifolds, rather than
2653: orbifolds, which makes the integration formulae easier to state.
2654: In fact, Theorems \ref{main} and \ref{ordinary}
2655: generalize easily to the orbifold case, as in \cite[\S 6]{Ma}.}
2656: Let $$M\mmod G = \muh^{-1}(\xi,0)/G\hs\hs\hs\text{ and }\hs\hs\hs
2657: M\mmod T = \mubh^{-1}(pr(\xi),0)/T$$
2658: be the hyperk\"ahler quotients of $M$ by $G$ and $T$, respectively.
2659: Because $\muh$ and $\mubh$ are circle-equivariant,
2660: the action of $S^1$ on $M$ descends to actions on the hyperk\"ahler quotients.
2661: Note that $M\mmod T$ also inherits an action of the Weyl group $W$ of $G$.
2662:
2663: \begin{example}\label{standard}
2664: The main example to keep in mind is $M=\cot$, where $\so$
2665: acts on $M$ by scalar multiplication on the fibers and the action
2666: of $G$ on $M$ is induced by a linear action of $G$ on $\Cn$,
2667: as in Chapter \ref{analogues}.
2668: \end{example}
2669:
2670: Consider the Kirwan maps
2671: $$\k_G:H_{\so\times G}^*(M)\to \hso(M\mmod G)\hs\hs\hs\text{ and }\hs\hs\hs
2672: \k_T:H_{\so\times T}^*(M)\to \hso(M\mmod T),$$
2673: induced by the inclusions of $\muh^{-1}(\xi,0)$ and $\mubh^{-1}(pr(\xi),0)$ into $M$,
2674: along with their rationalizations
2675: $$\hat\k_G:\widehat H_{\so\times G}^*(M)\to \hhso(M\mmod G)\hs\hs\hs\text{ and }\hs\hs\hs
2676: \hat\k_T:\widehat H_{\so\times T}^*(M)\to \hhso(M\mmod T).$$
2677: Let $$\rgt:\widehat{H}^*_{\so\times G}(M)\to\widehat{H}^*_{\so\times T}(M)^W$$
2678: be the standard isomorphism.
2679:
2680: % All of these maps are compatible with the module structures
2681: % of each of the rings over the ground ring
2682: % $H_{\so\times G}(pt) \cong (\operatorname{Sym}\tdu)^W\otimes\Q[x]$,
2683: % which includes into the rings $\widehat{H}_{\so\times G}(pt)$,
2684: % $H_{\so\times T}(pt)$, and $\widehat{H}_{\so\times T}(pt)$.
2685: Let $\Delta=\Delta^+\sqcup\Delta^-\subset\tdu$ be the set of roots of $G$.
2686: Let $$e = \Pd_{\a\in\Delta}\a(x-\a) \in (\operatorname{Sym}\tdu)^W\otimes\Q[x]
2687: \cong H_{\so\times G}(pt)\subs\widehat{H}_{\so\times G}(pt),$$
2688: and $$e' = \Pd_{\a\in\Delta^-}\a\cdot\Pd_{\a\in\Delta}(x-\a)
2689: \in\operatorname{Sym}\tdu\otimes\Q[x]\cong H_{\so\times T}(pt)
2690: \subs\widehat{H}_{\so\times T}(pt).$$
2691: The following two theorems are analogues of
2692: Theorems B and A of \cite{Ma}, adapted to circle compact
2693: hyperk\"ahler quotients. Our proofs follow closely those of Martin.
2694:
2695: \begin{theorem}\label{integration}
2696: Suppose that $M\mmod G$ and $M\mmod T$ are both circle compact.
2697: If $\ga\in\hhsg$, then
2698: $$\intxg \Kh_G(\ga) = \frac{1}{|W|}\intxt \Kh_T\circ\rgt(\ga)\cdot e.$$
2699: \end{theorem}
2700:
2701: \begin{theorem}\label{main}
2702: Suppose that $M\mmod G$ and $M\mmod T$ are both circle compact,
2703: and that the rationalized Kirwan map $\Kh_G$
2704: is surjective. Then
2705: $$\hhso(M\mmod G)\cong\frac{\hhso(M\mmod T)^W}{ann(e)}
2706: \cong\left(\frac{\hhso(M\mmod T)}{ann(e')}\right)^W.$$
2707: % and $$\hso(M\mmod G)\supseteq\Im(\K_G)\cong\frac{\hso(M\mmod T)^W}{ann(e)}
2708: % \cong\left(\frac{\hso(\M\mmod T)}{ann(e')}\right)^W,$$
2709: \end{theorem}
2710:
2711: \begin{proofintegration}
2712: Consider the following pair of maps:
2713: $$\begin{CD}
2714: \muh^{-1}(\xi,0)/T @>i>> \mubh^{-1}(pr(\xi),0)/T\cong M\mmod T\\
2715: @V\pi VV \\
2716: \muh^{-1}(\xi,0)/G\cong M\mmod G.
2717: \end{CD}$$
2718: Each of these spaces is a complex $S^1$-manifold with a compact, complex
2719: fixed point set, and therefore satisfies the hypotheses of Section
2720: \ref{thepush}.
2721: Let $$b = \Pd_{\a\in\Delta^+}\a\in H^*_{\so\times T}(pt)$$ be the product of the
2722: positive roots of $G$, which we will think of as an
2723: element of $\hhxt$. Martin shows that $\pi_*i^*b=|W|$,
2724: and that $i^*\circ\Kh_T\circ\rgt = \pi^*\Kh_G$ \cite{Ma}, hence
2725: we have
2726: \begin{eqnarray*}
2727: \intxg\Kh_G(\ga) &=& \frac{1}{|W|}\intxg\Kh_G(\ga)\cdot\pi_*i^*b\\
2728: &=& \frac{1}{|W|}\int_{\muh^{-1}(\xi,0)/T}\pi^*\Kh_G(\ga)\cdot i^*b\hspace{20pt}
2729: \text{by Definition~\ref{push}}\\
2730: &=& \frac{1}{|W|}\int_{\muh^{-1}(\xi,0)/T}i^*\circ\Kh_T\circ\rgt(\ga)\cdot i^*b\\
2731: &=& \frac{1}{|W|}\intxt \Kh_T\circ\rgt(\ga)\cdot b \cdot i_*(1)\hspace{20pt}
2732: \text{by Lemma~\ref{Npush}}.
2733: \end{eqnarray*}
2734: It remains to compute $i_*(1)\in\hhxt$.
2735: For $\a\in\Delta$, let $$L_{\a} = \mubh^{-1}((pr(\xi),0)\times_T\C_{\a}$$
2736: be the line bundle on $M\mmod T$ with $S^1$-equivariant
2737: Euler class $\a$.
2738: Similarly, let $L_x$ be the (topologically trivial)
2739: line bundle with $S^1$-equivariant
2740: Euler class $x$.
2741: Following the idea of \cite[1.2.1]{Ma}, we observe
2742: that the restriction of $\muh - (\xi,0)$ to
2743: $\mubh^{-1}(pr(\xi),0)$ defines
2744: an $\so\times T$-equivariant map $$s:\mubh^{-1}(pr(\xi),0)
2745: \to V\oplus V_{\C},$$ where $V = pr^{-1}(0)$
2746: and $V_{\C}=pr_{\C}^{-1}(0)$.
2747: This descends to an $S^1$-equivariant section of the
2748: bundle $$E = \mubh^{-1}(pr(\xi),0)\times_T \left(V\oplus V_{\C}\right)$$
2749: with zero locus $\muh^{-1}(\xi,0)/T$. The fact that $(\xi,0)$
2750: is a regular value implies that this section is generic,
2751: hence the equivariant Euler class $e(E)\in\hhxt$ is equal to $i_*(1)$.
2752:
2753: The vector space $V$ is isomorphic as a $T$-representation to
2754: $\bigoplus_{\a\in\Delta^-}\C_{\a}$,
2755: with $S^1$ acting trivially.
2756: Similarly, $V_{\C}$ is isomorphic to $V\otimes\C\cong V\oplus V^*$,
2757: with $S^1$ acting diagonally by scalars.
2758: Hence
2759: \begin{eqnarray*}
2760: E &\cong& \bigoplus_{\a\in\Delta^-}L_{\a}
2761: \oplus\bigoplus_{\a\in\Delta^-}\left(L_x\otimes L_{\a}\right)
2762: \oplus \left(L_x\otimes L_{-\a}\right)\\
2763: &\cong& \bigoplus_{\a\in\Delta^-}L_{\a}\oplus
2764: \bigoplus_{\a\in\Delta}L_x\otimes L_{-\a},
2765: \end{eqnarray*}
2766: and therefore $$i_*(1)=e(E) =
2767: \prod_{\a\in\Delta^-}\a\cdot\prod_{\a\in\Delta}(x-\a)=e'.$$
2768: Multiplying by $b$ we obtain $e$, and the theorem is proved.
2769: \end{proofintegration}
2770:
2771: % THIS IS STUPID; ANY RESTRICTION TO THE FIXED POINT SET
2772: % PROVIDES SUCH AN EXAMPLE.
2773: % Before proceeding with the proof of Theorem \ref{main} we present the
2774: % following example of Michael Thaddeus \cite{Th},
2775: % which shows that it is possible
2776: % for an inclusion of circle compact manifolds to induce a surjection
2777: % in rationalized $\so$-equivariant cohomology,
2778: % while not inducing one in non-rationalized $\so$-equivariant cohomology.
2779: %
2780: % \begin{example}\label{thaddeus}
2781: % Let $\so$ act on $\C P^1\times\C P^1$ by
2782: % $$((x:y),(u:v))\mapsto ((\lambda x :y),(u:v))$$ and on $\C P^3$ by
2783: % $$(z_1:z_2:z_3:z_4)\mapsto (\lambda z_1: \lambda z_2: z_3:z_4).$$
2784: % Then the Segr\'e embedding $i:\C P^1\times\C P^1\to \C P^3$ given by
2785: % $$i\big((x:y),(u:v)\big)=(xu:xv:yu:yv)$$
2786: % is $\so$-equivariant, and clearly induces an isomorphism
2787: % on the fixed point sets of the
2788: % $\so$ action. Therefore $i^*: \hhso(\C P^3)\to \hhso(\C P^1\times \C P^1)$
2789: % is surjective, and in fact an isomorphism, however
2790: % $i^*:\hso(\C P^3)\to \hso(\C P^1\times \C P^1)$ is only an
2791: % injection and therefore cannot be surjective.
2792: % \end{example}
2793:
2794: \begin{proofmain}
2795: Observe that the restriction
2796: of $\pi^*$ to the Weyl-invariant part
2797: $\hhso\left(\muh^{-1}(\xi,0)/T\right)^W$ is given by the composition of
2798: isomorphisms
2799: $$\hhso\left(\muh^{-1}(\xi,0)/T\right)^W
2800: \cong\widehat{H}^*_{\so\times T}\left(\muh^{-1}(\xi,0)\right)^W
2801: \cong\widehat{H}^*_{\so\times G}\left(\muh^{-1}(\xi,0)\right)
2802: \cong\hhxg,$$
2803: hence we may define
2804: $$i^*_W:= (\pi^*)^{-1}\circ i^*:
2805: \hhxt^W\to\hhso\left(\muh^{-1}(\xi,0)/T\right)^W.$$
2806: Furthermore, we have
2807: $\Kh_G = i^*_W\circ\Kh_T\circ\rgt$, hence
2808: $i^*_W$ is surjective.
2809: As in \cite[\S 3]{Ma},
2810: \begin{eqnarray*}
2811: i^*_W(a)=0 &\iff& \forall c\in\hhxt^W, \intxg i^*_W(c)\cdot i^*_W(a)=0\hspace{15pt}
2812: \text{by \ref{perfect} and surjectivity of $i^*_W$}\\
2813: &\iff& \forall c\in\hhxt^W, \intxt c\cdot a\cdot e = 0\hspace{15pt}
2814: \text{by Theorem~\ref{integration}}\\
2815: &\iff& \forall d\in\hhxt, \intxt d\cdot a\cdot e=0\hspace{15pt}
2816: \text{by using $W$ to average $d$}\\
2817: &\iff& a\cdot e=0\hspace{15pt}\text{by Lemma~\ref{perfect}},
2818: \end{eqnarray*}
2819: hence $\ker i^*_W = ann(e)$.
2820: By surjectivity of $i^*_W$,
2821: \begin{eqnarray*}
2822: \hhso(M\mmod G) &\cong&
2823: \frac{\hhxt^W}{\ker i^*_W}\cong\frac{\hhxt^W}{ann(e)}.
2824: % &\cong& \frac{\hhxt^W}{ann(e)}\cong\left(\frac{\hhxt}{ann(e')}\right)^W.
2825: \end{eqnarray*}
2826: By a second application of Lemma~\ref{perfect}, for any $a\in\hhxt$, we have
2827: \begin{eqnarray*}
2828: i^*(a) = 0 &\impl& \forall f\in\hhso(\muh^{-1}(\xi,0)/T),
2829: \int_{\muh^{-1}(\xi,0)/T}f\cdot i^*(a) = 0\\
2830: &\impl& \forall c\in\hhxt,
2831: \int_{\muh^{-1}(\xi,0)/T} i^*(c)\cdot i^*(a)=0\\
2832: &\impl& \forall c\in\hhxt,
2833: \intxt c\cdot a\cdot i_*(1)=0\hspace{15pt}\text{by Lemma~\ref{Npush}}\\
2834: &\impl& a\cdot e'= a\cdot i_*(1) = 0\hspace{15pt}\text{by Lemma~\ref{perfect}},
2835: \end{eqnarray*}
2836: hence $\ker i^* \subs ann(e')$.
2837: This gives us a natural surjection
2838: $$\frac{\hhxt^W}{ann(e)}=\frac{\hhxt^W}{\ker i^*_W}
2839: \cong \left(\frac{\hhxt}{\ker i^*}\right)^W
2840: \to \left(\frac{\hhxt}{ann(e')}\right)^W,$$
2841: which is also injective because $e'$ divides $e$.
2842: This completes the proof of Theorem \ref{main}.
2843: \end{proofmain}
2844:
2845: For the non-rationalized version of Theorem~\ref{main}, we make the additional
2846: assumption that $M\mmod G$ and $M\mmod T$ are equivariantly formal
2847: $\so$-manifolds, i.e. that $\hxg$ and $\hxt$ are free modules
2848: over $\hp$. Proposition \ref{formality} tells us that this
2849: is the case whenever the circle action
2850: is hamiltonian and its moment map is proper and bounded below.
2851:
2852: \begin{theorem}\label{ordinary}
2853: Suppose that $M\mmod G$ and $M\mmod T$ are equivariantly formal, circle compact, and
2854: that the rationalized Kirwan map $\Kh_G$ is surjective.
2855: % and that at least one of the ordinary Kirwan maps $\K_G$ or $\K_T$
2856: % is surjective as well.
2857: Then
2858: $$\hso(M\mmod G)\supseteq\Im(\K_G)\cong\frac{(\Im\K_T)^W}{ann(e)}
2859: \cong\left(\frac{\Im\K_T}{ann(e')}\right)^W.$$
2860: \end{theorem}
2861:
2862: \begin{remark}\label{notsobad}
2863: In the context of Example~\ref{standard} with $pr\circ\mu$ proper,
2864: $M\mmod G$ and $M\mmod T$ are both circle compact and equivariantly
2865: formal (Proposition \ref{formality}) and $\K_T$ is always surjective
2866: (Theorem \ref{hm}).
2867: \end{remark}
2868:
2869: \begin{proofordinary}
2870: Consider the following exact commutative diagram
2871: $$
2872: \xymatrix{
2873: 0 \ar[r] & A \ar[r]\ar[d] & \hxt^W \ar[r]^{i_W^*} \ar[d]
2874: & \hxg \ar[d]\\
2875: 0 \ar[r] & \widehat{A} \ar[r] & \hhxt^W \ar[r]^{i_W^*} & \hhxg.}
2876: $$
2877: Equivariant formality implies that the downward maps in the above
2878: diagram are inclusions, hence the map on top labeled $i^*_W$
2879: is simply the restriction of the map on the bottom
2880: to the subring $\hxt\subs\hhxt$.
2881: We therefore have $$A = \widehat{A}\cap\hxt^W = ann(e).$$
2882: Just as in the rationalized case, we have $\K_G = i_W^*\circ\K_T\circ\rgt$,
2883: hence
2884: $$\Im(\K_G)\cong i_W^*\left(\Im\K_T\circ\rgt\right)
2885: \cong \frac{(\Im\K_T)^W}{ann(e)}.$$
2886:
2887: Now consider the analogous diagram
2888: $$
2889: \xymatrix{
2890: 0 \ar[r] & B \ar[r]\ar[d] & \hxt \ar[r]^{\!\!\!\!\!\!\!\!\! i^*} \ar[d]
2891: & \hso(\muh^{-1}(\xi,0)/T) \ar[d]\\
2892: 0 \ar[r] & \widehat{B} \ar[r] & \hhxt \ar[r]^{\!\!\!\!\!\!\!\!\! i^*}
2893: & \hso(\muh^{-1}(\xi,0)/T).}
2894: $$
2895: Since we have not assumed that $\muh^{-1}(\xi,0)/T$ is equivariantly formal,
2896: we only know that the first two downward arrows are inclusions, and hence
2897: can only conclude that $B$ is contained in the annihilator of $e'$.
2898: Since $e'$ divides $e$, we have a series of natural surjections
2899: $$\frac{\imkt^W}{ann(e)}\cong
2900: \frac{\imkt^W}{A}\cong
2901: \left(\frac{\Im\K_T}{B}\right)^W\to
2902: \left(\frac{\Im\K_T}{ann(e')}\right)^W\to
2903: \left(\frac{\Im\K_T}{ann(e)}\right)^W.$$
2904: The composition of these maps is an isomorphism, hence so is each one.
2905: \end{proofordinary}
2906: \end{section}
2907: \end{chapter}
2908:
2909: \begin{chapter}{Hyperpolygon spaces}\label{hpspaces}
2910: A hyperpolygon space is the hyperk\"ahler analogue of a polygon space,
2911: which parameterizes $n$-sided polygons in $\R^3$ with fixed edge lengths.
2912: It is also an example of a quiver variety, introduced by Nakajima \cite{N1,N2},
2913: and since studied by many authors. In Section \ref{quiver} we give the basic
2914: constructions of quiver varieties and hyperpolygon spaces, and show that they
2915: satisfy all of the hypotheses of Chapter \ref{abelianization}.
2916: Section \ref{modcore} is devoted to understanding the components of the core
2917: of a hyperpolygon space; in particular, we show that they are all smooth
2918: (Theorem \ref{component}), and interpret them as moduli spaces of spatial polygons
2919: with certain properties (Theorem \ref{us}).
2920: Sections \ref{s1} and \ref{last} contain computations of the $\so$-equivariant
2921: cohomology rings of the hyperpolygon space as well as its core components,
2922: making use of the abelianization technique of Chapter \ref{abelianization}.
2923:
2924: This chapter is taken from \cite{HP2} and \cite{HP}. The reader is warned
2925: that our notation differs significantly from that of \cite{HP2}; most glaring
2926: is the fact that our spaces $\X$ and $\M$ correspond to the spaces
2927: $M$ and $X$ (respectively) in \cite{HP2}. This abrupt switch is necessary to
2928: conform with the conventions of Chapters \ref{analogues}-\ref{abelianization}.
2929:
2930: \begin{section}{Quiver varieties}\label{quiver}
2931: Let $Q$ be a quiver with vertex set $I$ and edge set $E\subs I\times I$,
2932: where $(i,j)\in E$ means that $Q$ has an arrow pointing from $i$ to $j$.
2933: We assume that
2934: $Q$ is connected and has no oriented cycles.
2935: Suppose given two collections of vector spaces $\{V_i\}$ and $\{W_i\}$,
2936: each indexed by $I$, and consider the affine space
2937: $$\Af = \bigoplus_{(i,j)\in E}\Hom(V_i,V_j)
2938: \,\oplus\,\bigoplus_{i\in I}\Hom(V_i,W_i).$$
2939: The group $U(V) = \prod_{i\in I}U(V_i)$ acts on $\Af$ by conjugation, and
2940: this action is hamiltonian. Given an element
2941: $$(B,J) = \bigoplus_{(i,j)\in E}B_{ij}\oplus\bigoplus_{i\in I}J_i$$
2942: of $\Af$,
2943: the $\mathfrak{u}(V_i)^*$ component of the moment map
2944: is $$\mu_i(B,J) = J_i^{\dagger}J_i+\sum_{(i,j)\in E}B_{ij}^{\dagger}B_{ij},$$
2945: where $\dagger$ denotes adjoint, and $\mathfrak{u}(V_i)^*$
2946: is identified with with the set of hermitian matrices via
2947: the trace form.
2948: Given a generic central element $\xi\in\mathfrak{u}(V)^*$,
2949: the K\"ahler quotient $\Af\,\mod_{\!\xi}U(V)$ parameterizes
2950: isomorphism classes of $\xi$-stable,
2951: framed representations
2952: of $Q$ of fixed dimension \cite{N2}. If $W_i=0$ for all $i$, then
2953: the diagonal circle $U(1)$ in the center of $U(V)$ acts trivially,
2954: and we instead quotient by
2955: $PU(V) = U(V)/U(1)$.
2956:
2957: Consider the hyperk\"ahler quotient $$\M = T^*\Af\,\mmod_{\!(\xi,0)}U(V),$$
2958: or, if $W_i=0$ for all $i$,
2959: $$\M = T^*\Af\,\mmod_{\!(\xi,0)}PU(V).$$
2960: As in Section~\ref{circleaction}, $\M$ has a natural action
2961: of $\gm$ induced from scalar
2962: multiplication on the fibers of $T^*\Af$.
2963: We now show that $M=T^*\Af$ satisfies
2964: the hypotheses of Theorems~\ref{main} and \ref{ordinary}.
2965:
2966: \begin{proposition}
2967: Let $T(V)\subs U(V)$ be a maximal torus, and let $pr:\mathfrak{u}(V)^*\to\mathfrak{t}(V)^*$
2968: be the natural projection.
2969: The moment maps $\mu=\Od_{i\in I}\mu_i:\Af\to\mathfrak{u}(V)^*$
2970: and $pr\circ\mu:\Af\to\mathfrak{t}(V)^*$ are each proper.
2971: \end{proposition}
2972:
2973: \begin{proof}
2974: To show that $\mu$ and $pr\circ\mu$ is proper, it suffices to find an element
2975: $t\in T(V)\subs U(V)$ such that the weights of the
2976: action of $t$ on $\Af$ are all strictly positive.
2977: Let $\lambda=\{\lambda_i\mid i\in I\}$ be a collection of integers,
2978: and let $t\in T(V)$ be the central element of $U(V)$ that acts
2979: on $V_i$ with weight $\lambda_i$ for all $i$.
2980: Then $t$ acts on $\Hom(V_i,V_j)$ with weight $\lambda_j-\lambda_i$,
2981: and on $\Hom(V_i,W_i)$ with weight $-\lambda_i$.
2982: Hence we have reduced the problem to showing that it is possible
2983: to choose $\lambda$ such that $\lambda_i<0$ for all $i\in I$ and
2984: $\lambda_i<\lambda_j$ for all $(i,j)\in E$.
2985:
2986: We proceed by induction on the order of $I$.
2987: Since $Q$ has no oriented cycles, there must exist a source $i\in I$;
2988: a vertex such that for all $j\in I$, $(j,i)\notin E$.
2989: Deleting $i$ gives a smaller (possibly disconnected) quiver with
2990: no oriented cycles, and therefore we may choose
2991: $\left\{\lambda_j\mid j\in I\smallsetminus\{i\}
2992: \right\}$ such that $\lambda_j<0$ for all $j\in I\smallsetminus\{i\}$
2993: and $\lambda_j<\lambda_k$ for all $(j,k)\in E$.
2994: We then choose $\lambda_i<\min\left\{\lambda_j\mid j\in I\smallsetminus\{i\}\right\}$,
2995: and we are done.
2996: \end{proof}
2997:
2998: \vspace{-\baselineskip}
2999: \begin{proposition}
3000: The rationalized Kirwan map
3001: $\Kh_{U(V)}:\widehat{H}^*_{\so\times U(V)}\Big(T^*\Af\Big)\to\hhso(\M)$
3002: is surjective.
3003: \end{proposition}
3004:
3005: \begin{proof}
3006: Nakajima \cite[\S 7.3]{N2} shows that there exist cohomology classes
3007: $a_i, b_i$ in the image of $\Kh_{U(V)}$
3008: such that $\dso = \sum\pi_1^*a_i\cdot\pi_2^*b_i$.
3009: (Nakajima uses a slightly different circle action, but
3010: his proof is easily adapted to the circle action that we have defined.)
3011: It follows from Proposition~\ref{basis} that
3012: the classes $\{b_i\}$ generate $\hhso(\M)$.
3013: \end{proof}
3014:
3015: \vspace{-\baselineskip}
3016: \begin{remark}\label{generation}
3017: This Proposition shows that the assumptions of
3018: Theorems~\ref{integration},~\ref{main},
3019: and~\ref{ordinary} are satisfied for Nakajima's quiver varieties.
3020: Thus integration in equivariant cohomology yields a
3021: description of the rationalized $\so$-equivariant cohomology, and also of
3022: the image of the non-rationalized Kirwan map $\K_G$. Therefore
3023: if we know that $\K_G$ is surjective for a particular quiver variety,
3024: then we have a concrete description of the ($\so$-equivariant)
3025: cohomology ring of that quiver variety.
3026: It is known that $\K_G$ is surjective
3027: for Hilbert schemes of $n$ points on an ALE space,
3028: so our theory applies and gives a description
3029: of the cohomology ring of these quiver varieties.
3030: More examples of quiver varieties with surjective Kirwan map,
3031: including hyperpolygon spaces,
3032: are discussed in Remark \ref{horn}.
3033: \end{remark}
3034:
3035: \begin{remark}
3036: Another interesting application of Proposition~\ref{basis} is
3037: to the moduli space ${\mathcal M}$ of stable rank $n$ and degree $1$
3038: Higgs bundles on a genus $g>1$ smooth projective algebraic curve $C$ (see \cite{H2}).
3039: It is an easy exercise to write down the cohomology class
3040: of the diagonal in ${\mathcal M}\times {\mathcal M}$
3041: as an expression in a certain set of tautological classes.
3042: Proposition~\ref{basis} implies that
3043: the rationalized $\so$-equivariant cohomology ring $\hhso({\mathcal M})$
3044: is generated by these classes.
3045: In fact the same result follows from the argument of \cite{HT1}.
3046: There ${\mathcal M}$ was embedded into a circle compact manifold
3047: ${\mathcal M}_\infty$, whose cohomology is the free algebra on the
3048: tautological classes. The argument in \cite{HT1} then goes by
3049: showing that the embedding of the $\so$-fixed point set of ${\mathcal M}$
3050: in that of ${\mathcal M}_\infty$ induces a surjection
3051: on cohomology. This already implies that $\hhso({\mathcal M}_\infty)$
3052: surjects onto $\hhso(\mathcal M)$. In \cite{HT1} it is shown
3053: that in the rank $2$ case this embedding also implies the surjection on
3054: non-rationalized cohomology, and then
3055: a companion paper \cite{HT2} describes the cohomology ring of ${\mathcal M}$ explicitly.
3056: However for higher rank
3057: this part of the argument of \cite{HT1} breaks down.
3058: Later Markman \cite{Mk} used similar diagonal arguments on
3059: certain compactifications of $\mathcal M$ and
3060: Hironaka's celebrated theorem on desingularization of algebraic
3061: varieties to deduce that the cohomology ring of ${\mathcal M}$
3062: is generated by tautological classes for all $n$.
3063: \end{remark}
3064:
3065: A {\em hyperpolygon space}, introduced in \cite{K2},
3066: is a quiver variety associated to the
3067: following quiver (Figure \ref{fig:quiver}),
3068: with $V_0 = \C^2$, $V_i=\C^1$ for $i\in\{1,\ldots,n\}$, and $W_i=0$ for all $i$.
3069: % It is so named because, for
3070: % $$\xi = \left(-\half\sum_{i=1}^n\xi_i;\hspace{2pt}\xi_1,\ldots,\xi_n\right)
3071: % \in\mathfrak{pu}(V)^*\subs\mathfrak{u}(2)^*\oplus\mathfrak{u}(1)^n,$$
3072: % the K\"ahler quotient $\Af\mod_{\!\xi}PU(V)\cong(\C^2)^n\mod_{\!\xi}PU(V)$
3073: % parameterizes $n$-sided polygons in $\R^3$
3074: % with edge lengths $\xi_1,\ldots,\xi_n$, up to rotation \cite{HK}.
3075: \begin{figure}[h]
3076: \centerline{\epsfig{figure=polyquiv.eps,height=4cm}}
3077: \caption{The quiver for a hyperpolygon space.}\label{fig:quiver}
3078: \end{figure}
3079:
3080: Let $$G := PU(V) = \Big(SU(2)\times U(1)^n\Big)\Big/\Z_2,$$
3081: and $$\GC := PGL(V) = \Big(\SL\times (\gm)^n\Big)\Big/\Z_2,$$
3082: where $\Zt$ acts by multiplying each factor by $-1$.
3083: We represent an element of $\Af\cong\ctn$ by an $n$-tuple
3084: of column vectors $$q = (q_1,\ldots,q_n).$$
3085: Following the conventions in \cite{K2}, we consider the {\em right}
3086: action of $G\subs\GC$ on $\Af$ given explicitly by
3087: $$q[\Theta;e_1,\ldots,e_n] = (\Theta^{-1}q_1e_1,\ldots,\Theta^{-1}q_ne_n).$$
3088: The compact group $G$ acts with moment map $\mu:\ctn\to\sutd\oplus\tnd$
3089: given by the equation $$\mu(q) = \sum_{i=1}^n (q_iq_i^*)_0 \oplus
3090: \left(\half|q_1|^2,\ldots,\half|q_n|^2\right),$$
3091: where $q_i^*$ denotes the conjugate transpose of $q_i$, $(q_iq_i^*)_0$
3092: denotes the
3093: traceless part of $q_iq_i^*$, and $\sutd$ is identified with
3094: $i\cdot\sut$ via the trace form.
3095: Given an $n$-tuple of real numbers $(\a_1,\ldots,\a_n)$,
3096: we define the {\em polygon space} $$\X(\a) := \ctn\mod_{\!\!\a}G,$$
3097: where $\a = 0\oplus (\a_1,\ldots,\a_n)\in\sutd\oplus\tnd$.
3098: If we break the reduction into two steps, reducing first by $U(1)^n$
3099: and then by $SU(2)$, we find that
3100: \begin{equation}\label{polygon}
3101: \X(\a)\cong\left\{(v_1,\ldots,v_n)\in (\R^3)^n\,\bigg|\,
3102: \|v_i\|=\a_i\text{ and }\sum v_i=0\right\}\bigg/SO(3)
3103: \end{equation}
3104: (see Remark \ref{interp} and the proof of Theorem \ref{us}).
3105: Here $\sutd$ is being identified with $\R^3$,
3106: and the coadjoint action of $SU(2)$
3107: on $\sutd$ is being replaced by the standard action of $SO(3)$ on $\R^3$ \cite{HK}.
3108: This space, therefore, may be thought of as the moduli space of $n$-sided polygons in $\R^3$,
3109: with fixed edge lengths $(\a_1,\ldots,\a_n)$, up to rotation.
3110: In particular, $\X(\a)$ is empty unless $\a_i\geq 0$ for all $i$.
3111:
3112: We call $\a$ {\em generic} if there does not exist a
3113: % map $\eps:\{1,\ldots,n\}\to\{\pm 1\}$
3114: subset $S\subs\{1,\ldots,n\}$
3115: such that $\sum_{i\in S}\a_i=\sum_{j\in S^c}\a_j$.
3116: Geometrically, this means that there is no element of $\X(\a)$ represented by a polygon that is
3117: contained in a single line in $\R^3$. If $\a$ is generic, then $\X(\a)$ is smooth \cite{HK}.
3118: Throughout this paper we will assume that $\a$ is generic, and that $\a_i> 0$ for all $i$.
3119:
3120: To define the hyperk\"ahler analogue of $\X(\a)$, we consider the induced action of $G$
3121: on $\cot$.
3122: % \cong\H^{2n}$, a $(2n)$-dimensional quaternionic "vector space".
3123: Explicitly, we write an element of $T^*\Af$ as $(p,q)$,
3124: where $q = (q_1,\ldots,q_n)$ is an $n$-tuple
3125: of column vectors
3126: and $p = (p_1,\ldots,p_n)$ an $n$-tuple
3127: of row vectors, and we put
3128: $$(p,q)[\Theta;e_1,\ldots,e_n] =
3129: \big((e_1^{-1}p_1\Theta,\ldots,e_n^{-1}p_n\Theta),(\Theta^{-1}q_1e_1,\ldots,\Theta^{-1}q_ne_n)\big).$$
3130: The action of $G$ on $T^*\Af$ is hyperhamiltonian
3131: with hyperk\"ahler moment map
3132: $$\mur\oplus\muc:\cot\to\Big(\sutd\oplus\tnd\Big)\oplus\Big(\slt^*\oplus
3133: (\mathfrak{u}(1)^n_{\C})^*\Big)$$
3134: given by the equations
3135: \begin{equation*}\label{mur}
3136: \mu_{\R}(p,q) = \frac{\sqrt{-1}}{2}\sum_{i=1}^n \left(q_i q_i^* - p_i^* p_i\right)_0
3137: \oplus\left(\half\left(|q_1|^2 - |p_1|^2\right), \ldots, \half\left(|q_n|^2 - |p_n|^2\right)\right)
3138: \end{equation*}
3139: and
3140: \begin{equation*}\label{muc}
3141: \mu_{\C}\left(p,q\right) = - \sum_{i=1}^n \left(q_i p_i\right)_0\oplus \left(\sqrt{-1} p_1 q_1,
3142: \ldots, \sqrt{-1} p_n q_n\right).
3143: \end{equation*}
3144: We then define the hyperpolygon space to be the hyperk\"ahler quotient
3145: $$\M(\a) := \cot\mmod_{(\a,0)}G =
3146: \Big(\mur^{-1}(\a)\cap\muc^{-1}(0)\Big)\Big/G,$$
3147: a smooth, noncompact hyperk\"ahler manifold of complex dimension $2(n-3)$.
3148: Recall that we also have
3149: $$\X(\a)\cong\left(\ctn\right)^{\ses}\!\big/G_{\C}
3150: \hspace{.7cm}\text{ and }\hspace{.7cm}
3151: \M(\a)\cong\muc^{-1}(0)^{\ses}\big/G_{\C},$$
3152: where $\ses$ means stable with respect to the weight $\a$ in the sense of
3153: geometric invariant theory (see Section \ref{reduction}).\footnote{Recall
3154: from Theorem \ref{hkhs} that the notions of stability and semistability
3155: agree for generic $\a$.}
3156: Nakajima gives a stability criterion for general quiver varieties \cite{N1,N2},
3157: which Konno interprets in the special case of hyperpolygon spaces.
3158: Call a subset $S\subs\{1,\ldots,n\}$ {\em short} if $\sum_{i\in S}\a_i < \sum_{j\in S^c}\a_j$,
3159: and call it {\em long} if its complement is short. (Assuming that $\a$ is generic is equivalent to assuming
3160: that every subset is either short or long.)
3161: Given a point $(p,q)\in\cot$ and a subset $S\subs\{1,\ldots,n\}$,
3162: we will say that $S$ is {\em straight}
3163: in $(p,q)$ if $q_i$ is proportional to $q_j$ for every $i,j\in S$.
3164: The terminology comes from K\"ahler polygon spaces, in which this condition
3165: is equivalent to asking that the vectors $v_i$ and $v_j$ be proportional
3166: over $\R_+$, or that the edges of lengths $\a_i$ and $\a_j$ (if they happen to be adjacent)
3167: line up to make a single edge of length $\a_i+\a_j$,
3168: as in Figure \ref{straight}.
3169: \begin{figure}[h]
3170: \begin{center}
3171: \psfrag{a1}{$\a_1$}
3172: \psfrag{a2}{$\a_2$}
3173: \psfrag{a3}{$\a_3$}
3174: \psfrag{a4}{$\a_4$}
3175: \psfrag{a5}{$\a_5$}
3176: \psfrag{a6}{$\a_6$}
3177: \psfrag{a7}{$\a_7$}
3178: \psfrag{a8}{$\a_8$}
3179: \includegraphics[height=40mm]{straight.eps}
3180: \caption{The subset $\{1,2,3\}$ is straight.}\label{straight}
3181: \end{center}\end{figure}
3182:
3183: \begin{theorem}\label{stability}{\em\cite[4.2]{K2}}
3184: Suppose that $\a$ is generic, and $\a_i> 0$ for all $i$.
3185: Then a point $(p,q)\in\cot$ is stable with respect to $\a$
3186: if and only if the following two conditions are satisfied:
3187: \begin{eqnarray*}
3188: &1)& q_i\neq 0 \text{ for all }i,\text{ and}\\
3189: &2)& \text{if $S$ is straight and $p_j=0$ for all $j\in S^c$, then $S$ is short.}
3190: \end{eqnarray*}
3191: \end{theorem}
3192: As in Chapter \ref{analogues}, we will use the notation $[p,q]$ to denote
3193: the $\GC$-equivalence class of a point $(p,q)\in\muc^{-1}(0)^{\ses}$,
3194: and $[p,q]_{\R}$ to denote
3195: the $G$-equivalence class of a point $(p,q)\in\mur^{-1}(\a)\cap\muc^{-1}(0)$.
3196: Recall that $\X$ sits inside of $\M$
3197: as the locus of points $[p,q]$ with $p=0$.
3198: This observation, along with Theorem \ref{stability}, allows us to recover
3199: the stability condition for the action of $G$ on $\ctn$.
3200: A point $q\in\ctn$ is stable if and only if $q_i\neq 0$ for all $i$,
3201: and no long subset is straight, as first shown in \cite{Kl}.
3202: The polygonally-minded reader is warned that in the hyperpolygon space $\M(\a)$,
3203: long subsets {\em can} be straight.
3204: \end{section}
3205:
3206: \begin{section}{Moduli theoretic interpretation of the core}\label{modcore}
3207: For the rest of the section we fix a generic
3208: $\a = 0\oplus (\a_1,\ldots,\a_n)\in\sutd\oplus\tnd$,
3209: with $\a_i>0$ for all $i$, and write $\M=\M(\a)$, $\X=\X(\a)$.
3210: % Consider the action of $\C^*$ on $\M$ given in the complex description
3211: % by $\la\cdot[p,q]=[\la p, q]$.
3212: % The circle $S^1\subs\C^*$ preserves
3213: % $\omega_{\R}$, and acts with moment map $\Phi:\M\to\R$ given in the symplectic
3214: % description by $\Phi\left([p,q]_{\R}\right) = \half\sum |p_i|^2$.
3215: Following Konno, we define
3216: $$\mathcal{S} = \big\{S\subs\{1,\ldots,n\}\hs\hs\big{|}\hs\hs S\text{ is short}\big\}$$
3217: and $$\spa = \big\{S\in\mathcal{S}\hs\hs\big{|}\hs\hs |S|\geq 2\big\}.$$
3218:
3219: \begin{theorem}\label{fixed}{\em\cite{K2}}
3220: The fixed point set
3221: $\displaystyle{\M^{\gm} = \M^{S^1} = \pola\cup\bigcup_{S\in\spa}\M_S,}$
3222: where
3223: $$\M_S = \big\{[p,q]\hs\hs\big{|}\hs\hs S\text{ and }S^c\text{ are each straight, and }
3224: p_j = 0\text{ for all }j\in S^c\big\}.$$ Furthermore, $\M_S$
3225: is diffeomorphic to $\C P^{|S|-2}$.
3226: \end{theorem}
3227:
3228: For all $S\in\spa$, let
3229: $U_S = U_{\polsa}$ be the piece of the core $\core\subs\M$
3230: defined in Section \ref{circleaction}.
3231: {\em A priori} we know only that $U_S$ is an irreducible, isotropic
3232: subvariety of dimension at most $n-3$ (Proposition \ref{coreprops}).
3233:
3234: \begin{theorem}\label{component}
3235: The core component $U_S$ is smooth of complex
3236: dimension $n-3$, and we have
3237: $$U_S = \big\{[p,q]\bigmid S\text{ is straight, and }
3238: p_j = 0\text{ for all }j\in S^c\big\}.$$
3239: \end{theorem}
3240:
3241: Before proving Theorem \ref{component},
3242: we describe the way in which the various components
3243: of the core fit together.
3244: For all $S\in\spa$, let $$\polsa = U_S\cap \pola =
3245: \big\{[0,q]\hs\hs\big{|}\hs\hs S\text{ is straight}\big\}.$$
3246: We call this space the {\em polygon subspace} of $\pola$
3247: corresponding to the short subset $S$.
3248: Note that $\polsa$ is itself a polygon space with $n-|S|+1$ edges, of lengths
3249: $\{\a_j\mid j\in S^c\}\cup\{\sum_S\a_i\}$. In particular, it is smooth.
3250: Now suppose given two short subsets $S,T\in\spa$,
3251: and consider the intersection $U_S\cap U_T$.
3252: \begin{itemize}
3253: \item If $S\cap T=\emptyset$, then $U_S\cap U_T = \polsa\cap X_T$, a polygon subspace
3254: both of $\polsa$ and of $X_T$.
3255: \item If $S\cap T\neq\emptyset$
3256: and $S\cup T$ is long, then $U_S\cap U_T=\emptyset$.
3257: \item If $S\cap T\neq\emptyset$ and
3258: $S\cup T$ is short,
3259: then $$U_S\cap U_T = \big\{[p,q]\bigmid S\cup T\text{ is straight, and }p_j=0\text{ for all }
3260: j\in (S\cap T)^c\big\}.$$
3261: This is a subvariety of $U_{S\cup T}$ given by taking the closure inside of $U_{S\cup T}$
3262: of a certain subbundle of the conormal bundle to $X_{S\cup T}\subs\pola$,
3263: defined by setting $p_j=0$ for all $j\in (S\cap T)^c\supseteq (S\cup T)^c$.
3264: \end{itemize}
3265: Each of these descriptions generalizes to higher intersections
3266: without modification.
3267:
3268: Finally, we compute the fixed point set $U_S^{\gm}$.
3269: If $[p,q]\in U_S^{\gm}$, then either $p=0$ and $[p,q]\in\polsa$,
3270: or $[p,q]\in \X_T$ for some $T\in\spa$. If $[p,q]\in \X_T$ then Theorem \ref{fixed}
3271: tells us that $T$ and $T^c$ are each straight, hence $S\subs T$ or $S\subs T^c$.
3272: Since $p\neq 0$, we must have $S\subs T$. Indeed, $U_S\cap \X_T$ is the linear subspace
3273: of $\X_T\cong \C P^{|T|-2}$ given by the condition $p_j=0$ for all $j\in T\smallsetminus S$.
3274: In particular, $U_S\cap \X_T$ is isomorphic to $\C P^{|S|-2}$ for any $T\supseteq S$.
3275:
3276: \begin{example}\label{ooh}
3277: Let $n=5$, $\a_1=\a_2=1$, and $\a_3=\a_4=\a_5=3$, and consider the short subset
3278: $S=\{1,2\}$. The fixed point set of $U_S$ consists of $\polsa\cong\C P^1$,
3279: and four points $\X_S$, $U_S\cap \X_{T_3}$,
3280: $U_S\cap \X_{T_4}$, and $U_S\cap \X_{T_5}$,
3281: where $T_j = \{1,2,j\}$ for $j=3,4,5.$
3282: For each $j$, $U_S\cap U_{T_j}$ is isomorphic to $\C P^1$,
3283: and touches $\polsa$ at the point $X_{T_j}$.
3284: In the following picture, an ellipse represents a copy of $\C P^1$ flowing between two
3285: fixed points, where the numbers or pairs of numbers
3286: indicate subsets that are straight on this $\C P^1$. (For example,
3287: $12,45$ means that $1$ and $2$ are straight,
3288: as are $4$ and $5$.)
3289: We will revisit this example at the end of Section \ref{last}.
3290: \begin{figure}[h]
3291: \begin{center}
3292: \psfrag{Xs}{$\M_S$}
3293: \psfrag{Xt3}{$\M_{T_3}$}
3294: \psfrag{Xt4}{$\M_{T_4}$}
3295: \psfrag{Xt5}{$\M_{T_5}$}
3296: \psfrag{Ms(a)}{$\X_{S}$}
3297: \psfrag{Mt3(a)}{$\X_{T_3}$}
3298: \psfrag{Mt4(a)}{$\X_{T_4}$}
3299: \psfrag{Mt5(a)}{$\X_{T_5}$}
3300: \includegraphics[height=40mm]{comp.eps}
3301: \caption{$U_S$, with $S=\{1,2\}$}
3302: \label{surface}
3303: \end{center}
3304: \end{figure}
3305: \end{example}
3306:
3307: \begin{componentproof}
3308: % The fact that $\dim U_S = \half\dim \M = n-3$ is a general property of
3309: % core components of quiver varieties \cite{N1}.
3310: % Thus, by a dimension count, it is enough to show that the set
3311: % $\big\{[p,q]\bigmid S\text{ is straight, and }
3312: % p_j = 0\text{ for all }j\in S^c\big\}$ is contained in $U_S$.
3313: %
3314: Consider a point $[p,q]\in \M$ with $S$ straight, and $p_j=0$ for all $j\in S^c$.
3315: By applying an element of $G$, we may assume that $q_i = \binom{1}{0}$ for all $i\in S$.
3316: Suppose further that there exists an $i\in S$ with $p_i\neq 0$, and that
3317: no strict superset of $S$ is straight. In other words, if
3318: $q_j = \binom{a_j}{b_j}$ for $j\in S^c$,
3319: suppose that $b_j\neq 0$.
3320: For $t\in\gm$, let $\Theta(t) =
3321: \footnotesize{\Big(\begin{array}{cc}
3322: t&0\\
3323: 0&t^{-1}
3324: \end{array}
3325: \Big)}$,
3326: let $e_i(t)=t$ for all $i\in S$,
3327: and let $e_j(t)=t^{-1}$ for all $j\in S^c$.
3328: Then for $i\in S$, we have $e_i(t)^{-1}p_i \Theta(t) = t^{-2} p_i$ and
3329: $\Theta(t)^{-1}q_i e_i = q_i$.
3330: For $j\in S^c$, we have $\Theta(t)^{-1}q_j e_j = \binom{t^{-2}a_j}{b_j}$.
3331: Hence
3332: \begin{eqnarray*}
3333: \lim_{t\to\infty}t\cdot[p,q] &=& \lim_{t\to\infty}t^2\cdot[p,q]\\
3334: &=& \lim_{t\to\infty}[t^2 p,q]\\
3335: &=& \lim_{t\to\infty}[t^2 e(t)^{-1}p \Theta(t), \Theta(t)^{-1}q e(t)]\\
3336: &=& [p,q'],
3337: \end{eqnarray*}
3338: where $q'_i = q_i$ for $i\in S$, and $q'_j = \binom{0}{b_j}$ for $j\in S^c$.
3339: Since we have assumed that $b_j\neq 0$ for all $j\in S^c$ and that $p_i\neq 0$
3340: for some $i\in S$, $(p,q')$ is stable, and hence defines an element of $\M_S$.
3341: Since $U_S$ is defined to be the closure of the set of elements that flow up to $\M_S$,
3342: it includes all $[p,q]$ with $S$ straight and $p_j=0$ for all $j\in S^c$.
3343: By dimension count, this containment is an equality, and
3344: we have $\dim U_S = n-3$.
3345:
3346: To see that $U_S$ is smooth, it is sufficient to show that $U_S$ is smooth
3347: at $[p,q]$ for all $[p,q]\in \M^{\gm}$.
3348: First suppose that $[p,q]\in \M_T$ for some $T\in\spa$ containing $S$.
3349: Suppose, without loss of generality, that
3350: $T=\{1,\ldots,l\}$ and $S=\{1,\ldots,m\}$ for some $l\leq m$.
3351: Konno computes an explicit local complex chart for $\M$ at the point $[p,q]$,
3352: with coordinates $\{z_i, w_i\mid 3\leq i\leq n-1\}$ \cite{K2}.
3353: With respect to these coordinates, a point $[p',q']$ has the property that
3354: $S$ is straight and $p'_j=0$ for all $j\in S^c$ if and only if
3355: $w_i=0$ for all $3\leq i\leq l$
3356: and $z_j=0$ for all $l+1 \leq j\leq n-1$.
3357: Hence $U_S$ is smooth at $[p,q]$.
3358:
3359: It remains only to show that $U_S$ is smooth at $\polsa = U_S\cap\pola$.
3360: Let $$E = \{(p,q)\mid S\text{ is straight, }
3361: p_j = 0\text{ for all }j\in S^c,\text{ and }\mu_{\C}(p,q)=0\},$$ and let
3362: $N = \{(p,q)\in E\mid p=0\}$. The natural projection from $E$ to $N$
3363: exhibits $E$ as a vector bundle over $N$, because the equation $\mu_{\C}(p,q)=0$
3364: is linear in $p$. We have $U_S = E\mod G = E^{\ses}/G$, and
3365: $\polsa = N\mod G = N^{\ses}/G$.
3366: The set $E\vert_{N^{\ses}}/G\subs E^{\ses}/G$ is
3367: an open neighborhood of $\polsa$ inside of $U_S$, and is isomorphic to a vector
3368: bundle over $\polsa$.
3369: Since $\polsa$ is a polygon space it is smooth, hence $U_S$ is smooth
3370: in a neighborhood of $\polsa$.
3371: \end{componentproof}
3372:
3373: \vspace{-\baselineskip}
3374: \begin{corollary}\label{compactification}
3375: $U_S$ is a compactification of the conormal
3376: bundle to $\polsa$ in $\pola$.
3377: \end{corollary}
3378:
3379: \begin{proof}
3380: Choose a point $[q,0]\in\polsa$, and a decomposition
3381: $$T_{[q,0]}\M = \nu_1\oplus \nu_2\oplus T_{[q,0]}\polsa\oplus E,$$
3382: where $\nu_1$ is the normal space to $\polsa$ inside of $U_S$,
3383: and $\nu_2$ is the normal space to $\polsa$ inside of $\pola$.
3384: Proposition \ref{coreprops} tells us that $U_S$ and $\pola$
3385: are both $\omega_{\C}$-lagrangian submanifolds of $\M$,
3386: hence $\omega_{\C}$ gives a perfect pairing between $T_{[q,0]}\polsa$
3387: and $E$. It follows that $\omega_{\C}$ also gives a perfect pairing between
3388: $\nu_1$ and $\nu_2$, and therefore that the normal bundle to
3389: $\polsa$ inside of $U_S$ is dual to the normal bundle of $\polsa$
3390: inside of $\pola$.
3391: % We must show that the normal bundle to $\polsa$ in $U_S$
3392: % is dual to the normal bundle to $\polsa$ in $\pola$.
3393: % We use only general facts about quiver varieties from \cite{N1},
3394: % and the additional information that $U_S$ is smooth,
3395: % from Theorem \ref{component}.
3396: % Consider a point $[0,q]\in\polsa$, and let $H_0$ and $H_1$ be the
3397: % $0$ and $1$ weight spaces of the $\C^*$ action on $T_{[0,q]}\M$.
3398: % The holomorphic symplectic
3399: % form $\omega_{\C}$ is being rotated by $\C^*$ with weight $1$ \cite[5.1]{N1},
3400: % hence it defines a perfect pairing between $H_0$ and $H_1$.
3401: % The fiber at $[0,q]$ of the normal bundle to $\polsa$ in $\pola$
3402: % is $H_0/T_{[0,q]}\polsa$, which is dual by $\omega_{\C}$ to the annihilator
3403: % of $T_{[0,q]}\polsa$ inside of $H_1$.
3404: % Since $U_S$ is $\C^*$-invariant, we may write
3405: % $$T_{[0,q]}U_S = T_{[0,q]}U_S\cap H_0\oplus T_{[0,q]}U_S\cap H_1
3406: % = T_{[0,q]}\polsa\oplus T_{[0,q]}U_S\cap H_1.$$
3407: % To prove Corollary \ref{compactification}, we must show that
3408: % $T_{[0,q]}U_S\cap H_1$ is equal to the annihilator of $T_{[0,q]}\polsa$.
3409: % The fact that $T_{[0,q]}U_S\cap H_1$
3410: % is contained in the annihilator of $T_{[0,q]}\polsa$ follows from the
3411: % fact that $U_S$ is lagrangian with respect to $\omega_{\C}$ \cite{N1}
3412: % (here we use
3413: % smoothness of $U_S$ at $[0,q]$).
3414: % Equality is then obtained by dimension count.
3415: \end{proof}
3416:
3417: \vspace{-\baselineskip}
3418: \begin{remark}
3419: This argument generalizes to the smooth intersection of any two
3420: lagrangian submanifolds of a symplectic manifold.
3421: \end{remark}
3422:
3423: We next describe $U_S$ in polygon-theoretic terms, as a certain moduli
3424: space of pairs of polygons in $\R^3$.
3425: % satisfying certain conditions,
3426: % up to rotation.
3427:
3428: \begin{theorem}\label{us}
3429: The core component $U_S$ is
3430: homeomorphic to the moduli space
3431: of $n+1$ vectors $$\{u_i, v_j, w\in\R^3\mid i\in S, j\in S^c\},$$
3432: taken up to rotation, satisfying the following conditions:
3433: \begin{eqnarray*}
3434: &1)&\hs\hs w + \sum_{j\in S^c}v_j = 0\\
3435: &2)&\hs\hs \sum_{i\in S}u_i = 0\\
3436: &3)&\hs\hs u_i\cdot w = 0\hs\hs\text{ for all }i\in S\\
3437: &4)&\hs\hs \|v_j\| = \a_j\hs\hs\text{ for all }j\in S^c\\
3438: &5)&\hs\hs \|w\| = \sum_{i\in S}\sqrt{\a_i^2 + \|u_i\|^2}.
3439: \end{eqnarray*}
3440: \end{theorem}
3441:
3442: \begin{remark}\label{interp}
3443: In more descriptive terms, a point in $U_S$ specifies two polygons in $\R^3$,
3444: as in Figure~\ref{fig:Uspoly}.
3445: \begin{figure}[h]
3446: \begin{center}
3447: \psfrag{w}{$w$}
3448: \psfrag{v1}{$v_1$}
3449: \psfrag{v2}{$v_2$}
3450: \psfrag{v3}{$v_3$}
3451: \psfrag{vnS}{$v_{n-|S|}$}
3452: \psfrag{u1}{$u_1$}
3453: \psfrag{u2}{$u_2$}
3454: \psfrag{uS}{$u_{|S|}$}
3455: \includegraphics[height=60mm]{Uspoly.eps}
3456: \caption{
3457: % The two polygons specified by a point in $U_S$. On the K\"ahler
3458: % polygon space at the minimum of $\Phi$ on $U_S$, the vectors $u_i$ are all
3459: % zero, and the vector $w$ corresponds to the $q_i$ for $i\in S$, which
3460: % are straight in $X_S$. At the maximum of $\Phi$ in $U_S$, the $v_i$
3461: % all lie on the same line as $w$, perpendicular to the $u_i$, making a
3462: % degenerate (linear) polygon. Then the only data left is given by the
3463: % second polygon specified by the $u_i$.
3464: An element of $U_S$, represented by a spatial polygon with a distinguished
3465: edge, and a planar polygon perpendicular to that edge.}
3466: \label{fig:Uspoly}
3467: \end{center}
3468: \end{figure}
3469: The first is the $n - |S| + 1$ sided polygon consisting of the vectors
3470: $\{v_j\mid j\in S^c\}$ and $w$. Each vector $v_j$ has length $\a_j$, and
3471: $w$ has a variable length, always greater than or equal to $\sum_{i\in S}\a_i$.
3472: This variable length is determined by the lengths of the edges in the second polygon,
3473: which consists of $|S|$ vectors $\{u_i\mid i\in S\}$,
3474: all contained in the plane perpendicular to $w$.
3475: Note that this description also applies to the K\"ahler polygon space $\X$
3476: by taking $S=\emptyset$.
3477:
3478: By setting $u_i = 0$ for all $i$ we get $\polsa$,
3479: the minimum of the Morse-Bott function $\Phi$ on $U_S$.
3480: On the other hand, consider the submanifold of $U_S$ obtained by imposing
3481: the extra condition that $\|w\| = \sum_{j\in S^c}\|v_j\|$.
3482: Then the first of the two polygons is forced to be linear,
3483: and we are left with $|S|$ vectors $\{u_i\}$ in the perpendicular plane
3484: satisfying a certain norm condition
3485: and adding to zero.
3486: Identifying this plane with $\C$ and dividing by the circle action rotating this plane,
3487: we obtain $\C P^{|S|-2}\cong \M_S$,
3488: the maximum of $\Phi$ on $U_S$.
3489: Other critical points of $\Phi$ occur whenever the first polygon is linear,
3490: which is possible for finitely many values of $\|w\|$.
3491: \end{remark}
3492:
3493: \begin{usproof}
3494: Suppose given a point $[p,q]_{\R}\in U_S$,
3495: and let
3496: $$u_i = q_ip_i + p_i^*q_i^*\hs\hs\text{ for all }i\in S,$$
3497: $$v_j = (q_j q_j^*)_0\hs\hs\text{ for all }j\in S^c,$$
3498: $$w = \sum_{i\in S}(q_iq_i^*)_0-(p_i^*p_i)_0.$$
3499: These vectors live in $i\cdot\sut\cong\sut^*\cong\R^3$, which is endowed with the
3500: metric
3501: $u\cdot v = \frac{1}{2}\operatorname{tr}uv$,
3502: invariant under the coadjoint action.
3503: With respect to this metric, we have the equalities
3504: $\|(qq^*)_0\| = \half |q|^2$ and $\|(p^*p)_0\| = \half |p|^2$,
3505: hence conditions (1), (2), and (4) are immediate consequences
3506: of the moment map equations.
3507:
3508: To verify condition (3), note that the vectors $\{q_i\mid i\in S\}$
3509: are all proportional over $\C$, which implies that the vectors $(q_iq_i^*)_0$
3510: are positive scalar multiples of each other.
3511: Furthermore, the moment map equation $p_iq_i=0$ implies that $(p_i^*p_i)_0$
3512: is a non-positive scalar multiple of $(q_iq_i^*)_0$, therefore
3513: $w = \sum(q_iq_i^*)_0-(p_i^*p_i)_0$ is proportional over $\R_+$ to
3514: $(q_iq_i^*)_0$ for any $i\in S$.
3515: Then $u_i\cdot w = \half\tr u_i w$ is a multiple of
3516: $$\tr u_i(q_iq_i^*)_0 = \tr u_iq_iq_i^* = \tr p_i^*q_i^*q_iq_i^*
3517: = |q_i|^2\tr p_i^*q_i^* = 0,$$
3518: where the first equality comes from the fact that
3519: $q_iq_i^*-(q_iq_i^*)_0$ is a scalar multiple of the identity,
3520: and $\tr u_i = 0$.
3521:
3522: To check condition (5), we first compute the norm of $u_i$:
3523: \begin{eqnarray*}
3524: \|u_i\|^2 &=& \half\tr u_i^2\\
3525: &=& \half\tr(q_ip_ip_i^*q_i^* + p_i^*q_i^*q_ip_i)\\
3526: &=& |q_i|^2|p_i|^2\\
3527: &=& |q_i|^2 (|q_i|^2 - 2\a_i).
3528: \end{eqnarray*}
3529: Since all of the vectors
3530: $\{(q_iq_i^*)_0 -(p_i^*p_i)_0\mid i\in S\}$ point in the same direction,
3531: we have $$\|w\| = \sum_{i\in S}\|(q_iq_i^*)_0\| + \|(p_i^*p_i)_0\|
3532: = \sum_{i\in S} \half|q_i|^2 + \half|p_i|^2
3533: = \sum_{i\in S} |q_i|^2 - \a_i
3534: = \sum_{i\in S} \sqrt{\a_i^2 + \|u_i\|^2}.$$
3535:
3536: We have defined a continuous map from $U_S$ to the moduli space of
3537: vectors $\{u_i,v_j,w\}$ satisfying conditions (1)-(5),
3538: and we claim that this map is a homeomorphism.
3539: Since the source of this map is compact and the target is Hausdorff,
3540: it is sufficient to show that the map is bijective.
3541:
3542: Suppose given a collection of vectors $\{u_i,v_j,w\}\subs\sut$
3543: satisfying conditions (1)-(5). Using the adjoint action of $SU(2)$,
3544: we may assume that $w$ is a positive scalar multiple of
3545: $\footnotesize{\Big(\begin{array}{cc}
3546: 1&0\\
3547: 0& -1
3548: \end{array}
3549: \Big)}$.
3550: By condition (3), this implies that for all $i\in S$, there exists $t_i\in\C$
3551: with $u_i =
3552: \footnotesize{\Big(\begin{array}{cc}
3553: 0& t_i\\
3554: \bar{t_i}& 0
3555: \end{array}
3556: \Big)}$.
3557: For $j\in S^c$, we choose $q_j\in\C^2$ with $(q_jq_j^*)_0=v_j$,
3558: and observe that $q_j$ is unique up to the action of $U(1)^n$.
3559: We know that for all $i\in S$, $(q_iq_i^*)_0$ must be a positive multiple of $w$,
3560: hence there exist $a_i,b_i\in\C$ such that
3561: $$q_i=\binom{a_i}{0}\hs\text{ and }\hs p_i = (0\hs b_i)$$
3562: for all $i\in S$.
3563: In order to have $u_i = q_ip_i + p_i^*q_i^*$ and $\half|q_i|^2-\half|p_i|^2=\a_i$,
3564: we must have $$a_i b_i = t_i\hs\text{ and }\hs
3565: \half|a_i|^2-\half|b_i|^2=\a_i.$$
3566: These equations uniquely define $a_i$ and $b_i$ up to the action of $U(1)^n$.
3567: It follows from conditions (1)-(5) that
3568: $(p,q)\in\mur^{-1}(\a)\cap\muc^{-1}(0)$
3569: and that $w=\sum_{i\in S}(q_iq_i^*)_0-(p_i^*p_i)_0$.
3570: This shows that our map is bijective, and thus completes the proof of Theorem \ref{us}.
3571: \end{usproof}
3572:
3573: \vspace{-\baselineskip}
3574: \begin{remark}
3575: Suppose that $S$ has only two elements; without loss of generality
3576: we will assume that $S=\{1,2\}$. Then forgetting $u_1$ and $u_2$
3577: gives a diffeomorphism from $U_S$ to the ``vertical polygon space''
3578: $V\!P(\a_3,\ldots,\a_n,\a_1+\a_2)$
3579: defined in \cite{HK}, shown to be diffeomorphic to a toric variety.
3580: More generally with $S=\{1,\ldots,k\}$, given any two-element subset $T\subs S$, the subvariety
3581: of $U_S$ given by the equations $u_i=0$ for all $i\in S\smallsetminus T$
3582: is diffeomorphic to $V\!P(\a_{k+1},\ldots,\a_n,\sum_T\a_i)$.
3583: \end{remark}
3584: \end{section}
3585:
3586: \begin{section}{Cohomology rings}\label{s1}
3587: In this section we use Theorem \ref{ordinary} to compute the circle-equivariant
3588: cohomology of a hyperpolygon space $\M$, thus reproducing (by different means)
3589: the results of \cite[\S 3]{HP2}.
3590: Recall that we have $$\M = T^*\ctn\mmod G,$$ where $G$ is a quotient
3591: of $U(1)^n\times SU(2)$ by $\Zt$.
3592: We will simplify our computations by dividing first by
3593: the torus $U(1)^n$.
3594: We have
3595: \begin{eqnarray*}
3596: \M &=& \left(T^*\C^{2n}\right)\bigmmod G\\
3597: &\cong&\left(\left(T^*\C^2\right)^n\bigmmod U(1)^n\right)\bigmmod SU(2)\\
3598: &\cong& \prod_{i=1}^nT^*\C P^1\bigmmod SU(2),
3599: \end{eqnarray*}
3600: where the action of $SU(2)$ on each copy of $T^*\C P^1$
3601: is induced by the rotation action on $\C P^1\cong S^2$.
3602:
3603: \begin{proposition}\label{hpsurj}
3604: The non-rationalized Kirwan map
3605: $\K_{U(V)}:H^*_{\so\times U(V)}(T^*\C^{2n})\to\hso(\M)$ is surjective.
3606: \end{proposition}
3607:
3608: \begin{proof}
3609: The map $\K_{U(V)}$ factors as a composition
3610: $$H^*_{\so\times U(V)}(T^*\C^{2n})
3611: \to H^*_{\so\times SU(2)}\left(\prod_{i=1}^nT^*\C P^1\right)
3612: \overset{\K_{SU(2)}}\Longrightarrow \hso(\M),$$
3613: where the first map is the Kirwan map for a toric hyperk\"ahler variety,
3614: and therefore surjective by \cite{HP1}. Hence it suffices to show that
3615: $\K_{SU(2)}$ is surjective.
3616:
3617: The level set $\muc^{-1}(0)$ for the action
3618: of $SU(2)$ on $\prod_{i=1}^nT^*\C P^1$ is a subbundle of the cotangent bundle,
3619: given by requiring the $n$ cotangent vectors to add to zero after being restricted
3620: to the diagonal $\C P^1$. In particular this set is smooth,
3621: and its $\so\times SU(2)$-equivariant cohomology ring is canonically
3622: isomorphic to that of $\prod_{i=1}^nT^*\C P^1$.
3623: Hence $\K_{SU(2)}$ factors as
3624: $$H^*_{\so\times SU(2)}\left(\prod_{i=1}^nT^*\C P^1\right)
3625: \cong H^*_{\so\times SU(2)}\Big(\muc^{-1}(0)\Big)\to
3626: \hso\Big(\muc^{-1}(0)\bigmod SU(2)\Big)
3627: \cong \hso(\M),$$ where the map in the middle is the K\"ahler
3628: Kirwan map.
3629: Surjectivity of this map follows from the following more general lemma,
3630: applied to the manifold $\muc^{-1}(0)$.
3631:
3632: \begin{lemma}\label{cut}
3633: Suppose given a hamiltonian action of $\so\times G$ on a symplectic
3634: manifold $M$, such that the $\so$ component of the moment map is proper
3635: and bounded below with finitely many critical values.
3636: Then the K\"ahler Kirwan map
3637: $\K:H^*_{\so\times G}(M)\to\hso(M\mod G)$ is surjective.
3638: \end{lemma}
3639:
3640: \begin{proof}
3641: Extend the action of $S^1$ to an action on $M\times\C$ by letting
3642: $S^1$ act only on the left-hand factor. On the other hand,
3643: consider a second copy of the circle, which we will call $\mathbb T$ to avoid confusion,
3644: acting diagonally on $M\times\C$.
3645: Choose $r\in \operatorname{Lie}(\mathbb T)^*\cong\R$ greater than
3646: the largest critical value of the $\mathbb T$-moment map, and consider
3647: the space $$Cut(M\mod G):=\left(M\times\C\right)\mod_{\! r} \mathbb T\times G
3648: \cong\big((M\mod G)\times\C\big)\mod_{\! r} \mathbb T.$$
3649: This space, which is called the {\em symplectic cut} of $M\mod G$ \cite{Le},
3650: is an $S^1$-equivariant (orbifold) compactification of $M\mod G$.
3651: We then have a commutative diagram
3652: $$\begin{CD}
3653: H^*_{\so\times G\times \mathbb T}(M\times\C) @>>> H^*_{\so\times G}(M)\\
3654: @VVV @VV\K V\\
3655: \hso(Cut(M\mod G)) @>>> \hso(M\mod G).
3656: \end{CD}$$
3657: The vertical map on the left is surjective because the $G\times \mathbb T$ moment map is proper,
3658: and the map on the bottom is surjective because the long
3659: exact sequence in cohomology for $M\mod G\subs Cut(M\mod G)$
3660: splits naturally, hence $\K$ is surjective as well.
3661: \end{proof}
3662:
3663: \noindent By applying Lemma~\ref{cut} to $M = \muc^{-1}(0)$,
3664: this completes the proof of Proposition~\ref{hpsurj}.
3665: \end{proof}
3666:
3667: \vspace{-\baselineskip}
3668: \begin{remark}\label{horn}
3669: The argument in Proposition \ref{hpsurj} generalizes immediately
3670: to show that the hyperk\"ahler Kirwan map for the quotient
3671: $$\left(\prod_{i=1}^nT^*Flag(\C^k)\right)\bigmmod SU(k)$$
3672: is surjective. This is itself a quiver variety, and like the
3673: hyperpolygon space, it has a moduli theoretic interpretation.
3674: The K\"ahler quotient
3675: $$\left(\prod_{i=1}^nFlag(\C^k)\right)\bigmod SU(k)$$
3676: is isomorphic to the space of $n$-tuples of $k\times k$
3677: hermitian matrices with fixed eigenvalues adding to zero, modulo
3678: conjugation.
3679: This space has been studied by many authors. The classical
3680: problem, due to Horn, of determining the values of
3681: the moment map for which it is
3682: nonempty, has only recently been solved \cite{KT}. For a survey, see \cite{Fu}.
3683: \end{remark}
3684:
3685: To compute the kernel of the hyperk\"ahler Kirwan map for the
3686: hyperpolygon space, we first need to study the abelian quotient
3687: $$\NN := \prod_{i=1}^nT^*\C P^1\bigmmod T,$$ where $T\cong U(1)\subs SU(2)$ is a
3688: maximal torus.\footnote{This is the hyperk\"ahler analogue of the {\em abelian
3689: polygon space} from \cite{HK}.}
3690: The space $\prod_{i=1}^nT^*\C P^1$ is a hypertoric variety
3691: given by an arrangement of $2n$ hyperplanes in $\R^n$,
3692: where the $(2i-1)^{\text{st}}$ and $(2i)^{\text{th}}$ hyperplanes
3693: are given by the equations $x_i = \pm\xi_i$.
3694: Taking the hyperk\"ahler quotient by $T$ corresponds on the level of arrangements
3695: to restricting this arrangement to the hyperplane
3696: $\{x\in\R^n\mid\sum x_i = 0\}$.
3697: By Theorem \ref{hsm}, we have
3698: $$\hso\left(\NN\right)\cong
3699: \Q[a_1,b_1,\ldots,a_n,b_n,\delta,x]\Big/
3700: \Big\la a_i-b_i-\delta, \hs a_ib_i\hs\Big{|}\hs i\leq n\Big\ra+
3701: \Big\la A_S, B_S\hs\Big{|}\hs S\text{ short}\Big\ra,
3702: $$
3703: % where $$\mathcal{I} = \left\la (a_i-b_i)-(a_j-b_j)\hs\hs\Big{|}\hs\hs i,j\leq n\right\ra,$$
3704: % $$\mathcal{M} = \left\la \prod_{i\in S}(x-a_i)\prod_{j\in S^c}b_j,\hs\hs
3705: % \prod_{i\in S}(x-b_i)\prod_{j\in S^c}a_j\hs\hs\Big{|}\hs\hs S\text{ short}\right\ra,$$
3706: % and $$\mathcal{K2}=\left\la c_id_i\mid i\leq n\right\la.$$
3707: where
3708: \begin{equation*}\label{asbs}
3709: A_S= \prod_{i\in S}(x-a_i)\prod_{j\in S^c}b_j\hs\hs
3710: \text{ and }\hs\hs B_S = \prod_{i\in S}(x-b_i)\prod_{j\in S^c}a_j.
3711: \end{equation*}
3712: Here $\delta$ is the image in $\hso\left(\NN\right)$
3713: of the unique positive root of $SU(2)$.
3714: The Weyl group $W$ of $SU(2)$, isomorphic to $\Z/2\Z$, acts on this ring by fixing $x$
3715: and switching $a_i$ and $b_i$ for all $i$.
3716: Let $c_i = a_i+b_i$, and let $C_S=A_S+B_S$.
3717: Let $$P=\Q[c_1,\ldots,c_n,\delta,x]\Big/
3718: \Big\la c_i^2-\delta^2\hs\Big{|}\hs i\leq n\Big\ra$$ and
3719: $$Q=P^W=\Q[c_1,\ldots,c_n,\delta^2,x]\Big/
3720: \Big\la c_i^2-\delta^2\hs\Big{|}\hs i\leq n\Big\ra.$$
3721: Let $$\I=\Big\la A_S, B_S\hs\Big{|}\hs S\text{ short}\Big\ra\subs P
3722: \hspace{15pt}\text{and}\hspace{15pt}
3723: \J=\Big\la C_S\hs\Big{|}\hs S\text{ short}\Big\ra\subs Q,$$
3724: so that $$\hso(\NN)\cong P/\I
3725: \hspace{15pt}\text{and}\hspace{15pt}
3726: \hso(\NN)^W\cong Q/\J.$$
3727: Note that all odd powers of $\delta$ in the expression for $C_S=A_S+B_S$ cancel out.
3728: Then by Theorem~\ref{ordinary} and Remark~\ref{notsobad},
3729: \begin{eqnarray*}
3730: \hso(\M)&\cong& \frac{\hso\left(\NN\right)^W}{ann(e)}
3731: \cong \frac{Q}{(e:\J)},
3732: \end{eqnarray*}
3733: where $e=\delta^2(x^2-\delta^2)$, and $(e:\J)$ is the ideal of elements
3734: of $Q$ whose product with $e$ lies in $\J$.
3735:
3736: If $S$ is a nonempty short subset, let
3737: $m_S$ be the smallest element of $S$, $n_S$ the smallest
3738: element of $S^c$, and $$D_S = \prod_{m_S\neq i\in S}(c_i-x)
3739: \,\,\cdot\!\prod_{n_S\neq j\in S^c}(c_{n_S}+c_j)\in Q.$$
3740:
3741: \begin{theorem}\label{hp}
3742: The circle-equivariant cohomology ring of the hyperpolygon space
3743: $\M$ is isomorphic to\footnote{The class
3744: denoted by $c_i$ in \cite{HP2} differs from our $c_i$ by a sign, hence
3745: to recover the presentation of \cite{HP2} we must replace $c_i-x$
3746: with $c_i+x$ in the expression for $D_S$.}
3747: $$Q\big/\big\la
3748: D_S\mid \emptyset\neq S\text{ short}\big\ra.$$
3749: \end{theorem}
3750:
3751: \begin{proof}
3752: We begin by proving that $e\cdot D_S\in\mathcal{J}$ for all nonempty
3753: short subsets $S\subs\{1,\ldots,n\}$. We will in fact prove the slightly
3754: stronger statement
3755: $$e\cdot D_S\in\Big\la C_T\hs\Big{|}\hs T\subs S\text{ short}\Big\ra
3756: \subs\mathcal{J},$$
3757: proceeding by induction on $|S|$.
3758: We will assume, without loss of generality, that $n\in S$.
3759: The base case occurs when $S=\{n\}$, and in this case
3760: we observe that
3761: $$e\cdot D_S = 2^{n-3}\cdot(x+c_n)\cdot\Big((2x-c_n)\cdot C_{\emptyset} - c_n\cdot C_S\Big).$$
3762: We now proceed to the inductive step, assuming that the proposition
3763: is proved for all short subsets of size less than $|S|$, and all values of $n$.
3764: For all $T\subs S\smallsetminus\{n\}$, we have
3765: $$\half\Big(C_T-C_{T\cup\{n\}}\Big)=(c_n-x)\cdot C'_T,$$
3766: where $C'_T$ is the polynomial in the variables $\{c_1,\ldots,c_{n-1},\delta^2\}$
3767: corresponding to the short subset $T\subs\{1,\ldots,n-1\}$.
3768: Since $S\smallsetminus\{n\}$ is a short subset of $\{1,\ldots,n-1\}$
3769: of size strictly smaller than $S$,
3770: our inductive hypothesis tells us that $e\cdot D_S/(c_n-x)$
3771: can be written as a linear combination of polynomials $C'_T$, where the coefficients
3772: are quadratic polynomials in $\{c_1,\ldots,c_{n-1},\delta^2\}$.
3773: Replacing $C'_T$ with $\half\left(C_T-C_{T\cup\{n\}}\right)=(c_n-x)\cdot C'_T$,
3774: we have expressed $e\cdot D_S$ in terms of the appropriate polynomials.
3775: This completes the induction.
3776:
3777: Suppose that $F\in Q$ is an element of degree less than $n-2$
3778: such that $e\cdot F\in\J$. By the second isomorphism
3779: of Theorem~\ref{ordinary}, this implies that $e'\cdot F\in\I\subs P$,
3780: where $e'=\delta(x^2-\delta^2)$.
3781: Consider the quotient ring $R$ of $P$ obtained by setting
3782: $a_i^2=b_i^2=x=0$ for all $i$. (Recall that $a_i=\half(c_i+\delta)$
3783: and $b_i=\half(c_i-\delta)$.) Then element $e'$ maps to zero in $R$,
3784: while the generators $\{A_S, B_S\}$ of $\I$ descend to a basis
3785: for the $n^{\text{th}}$ degree part of $R$. This means that we
3786: must have $e'\cdot F=0\in P$.
3787: Using the additive basis for $P$ consisting of monomials
3788: that are squarefree in the variables $c_1,\ldots,c_n$,
3789: it is easy to check that $e'$ is not a zero divisor in $P$,
3790: and therefore that $F=0$.
3791:
3792: Finally, we must show that $\{D_S\mid \emptyset\neq S\text{ short}\}$
3793: generates all elements of $(e:\J)$ of degree at least $n-2$.
3794: We obtain this fact from the following technical lemma, the proof of which we defer
3795: until the end of the section.
3796:
3797: \begin{lemma}\label{tech}
3798: The set $\{D_S\mid \emptyset\neq S\text{ short}\}$
3799: descends to a basis for the degree $n-2$
3800: part of the quotient ring $Q/\la x\ra$.
3801: \end{lemma}
3802:
3803: Let $F$ be an element of minimal degree $k\geq n-2$ that is in
3804: $(e:\J)$ but not $\la D_S\mid \emptyset\neq S\text{ short}\ra$.
3805: By Lemma \ref{tech}, $F$
3806: differs from an element of
3807: $\la D_S\mid \emptyset\neq S\text{ short}\ra$ by $x\cdot F'$ for some $F'$
3808: of degree $k-1$. By equivariant formality of $\hso(\M)$,
3809: $$x\cdot F' = F\in (e:\J)\impl F'\in (e:\J),$$
3810: which contradicts the minimality of $k=\operatorname{deg}F$.
3811: Hence $\la D_S\mid \emptyset\neq S\text{ short}\ra=(e:\J)$, and the proposition
3812: is proved.
3813: \end{proof}
3814:
3815: \vspace{-\baselineskip}
3816: \begin{corollary}\label{konno}
3817: The ordinary cohomology ring $H^*(\M)$ is isomorphic to
3818: $$\Q[c_1,\ldots,c_n]\Big/\big\la c_i^2-c_j^2\mid i,j\leq n\big\ra
3819: +\la\text{all monomials of degree $n-2$}\ra.$$
3820: \end{corollary}
3821:
3822: \begin{proof}
3823: This follows from the fact that $H^*(\M)\cong \hso(\M)/\la x\ra$
3824: for any equivariantly formal space $M$, and the observation in \cite{HP2}
3825: that $\{D_S\mid \emptyset\neq S\text{ short}\}$ descends to a basis for the degree $n-2$
3826: part of $Q/\la x\ra$.
3827: \end{proof}
3828:
3829: \begin{prooftech}
3830: Let $\bk = \half(c_1+\ck)$ for all $k$, so that $c_k = 2\bk-d_1$.
3831: The relations $c_k^2 = c_1^2$ translate to $d_k^2 = d_1d_k$ for all $k$,
3832: and we have
3833: $$Q/\la x\ra = \Q[d_1,\ldots,d_n]\Big/\big< d_k^2-d_1d_k\hs\big{|}\hs k\in\twn\big>.$$
3834: For all short subsets $S$, put
3835: $$\Sb = S\smallsetminus\{m_S\}\hspace{15pt}\text{and}\hspace{15pt}
3836: \Lb = S^c\smallsetminus\{n_S\},$$
3837: and let
3838: $$v_S = (-1)^n\pjl (\bj+\bns-d_1)\times\pis(2\bi-d_1).$$
3839: For all $A\subs\twn$, let
3840: $$\ba = (-1)^{|A|} d_1^{n-2-|A|} \prod_{k\in A} d_k$$ for all $A \subsetneq \twn$.
3841: Then $\{d_A\}$
3842: is a basis for
3843: the $(n-2)^{\text{nd}}$ graded piece of $Q/\la x\ra$,
3844: and $v_S$ is equal to $(-1)^n\cdot 2^{-|\Lb|}$ times the image of $D_S$ in $Q/\la x\ra$.
3845: Hence our the statement of Lemma \ref{tech} is that for all $A$, $\ba$ may be
3846: expressed as a linear combination of the elements $\{v_S\mid S\in\sa\}$.
3847:
3848: \begin{convention}
3849: The notation $S^c$ refers to the complement of $S$ inside of the set $\otn$,
3850: while the notation $A^c$ refers to the complement of $A$ inside of the set $\twn$.
3851: \end{convention}
3852:
3853: \begin{claim}\label{vs}
3854: We have the following expression for $v_S$ in terms of the basis $\{d_A\}$:
3855: $$\vs =
3856: \begin{cases}
3857: \displaystyle{\sum_{\ss{\Lb\hs\subseteq A \\ m_S\notin A}}} \hs 2^{|\asb|} \hs\ba
3858: & \text{if}\hspace{10pt}1 \in S^c;\vspace{.5cm}\\
3859: \displaystyle{\sum_{S^c \nsubseteq A}} \hs 2^{|\asb|}
3860: \hs\ba & \text{if}\hspace{10pt}1 \in S.\\
3861: \end{cases}
3862: $$
3863: \end{claim}
3864:
3865: \begin{proof}
3866: Any degree $n-2$ monomial
3867: in $d_1,\ldots,d_n$ is equal to $(-1)^{|A|}d_A$, where $A$ is the set of $k>1$ such that
3868: $d_k$ appears in the monomial.
3869: Expanding $\vs$, we need to count (with sign) the occurrence of $d_A$ for each $A$.
3870: In most cases we find that there is no cancellation, and the claim is straightforward.
3871: The most difficult case occurs when
3872: $1\in S$ (therefore $n_S=1$) and $\ms\in A$;
3873: in this case the number of times (with multiplicity) that $d_A$
3874: occurs in $\vs$ is
3875: \begin{eqnarray*}
3876: && (-1)^n(-1)^{|A|}(-1)^{|\acsb|}\hs 2^{|\asb|}\sum_{E\subsetneq\Ac\cap\Lb}(-1)^{|E|}
3877: \vspace{.3cm}\\
3878: \vspace{.3cm}&=& (-1)^n(-1)^{|A|}(-1)^{|\acsb|}\hs 2^{|\asb|}
3879: \left((1-1)^{|\Ac\cap\Lb|}-(-1)^{|\Ac\cap\Lb|}\right)\vspace{.3cm}\\
3880: % &=& (-1)^n(-1)^{|A|}(-1)^{|\acsb|}\hs 2^{|\asb|}(-1)^{|\Ac\cap\Lb|+1}\\
3881: \vspace{.3cm}&=& (-1)^{n+|A|+|\acsb|+|\Ac\cap\Lb|+1}\hs 2^{|\asb|}\\
3882: % &=& (-1)^{n+|A|+|\Ac|+1}\hs 2^{|\asb|}\\
3883: \vspace{.3cm}&=& (-1)^{2n}\hs 2^{|\asb|}\\
3884: \vspace{.3cm}&=& \hs 2^{|\asb|}.
3885: \end{eqnarray*}
3886: We leave the remaining cases to be checked by the reader.
3887: \end{proof}
3888:
3889: \vspace{-\baselineskip}
3890: \begin{claim}\label{ws}
3891: Suppose that $1\in S$.
3892: Let $S_0 = S$, and for $1\leq k\leq |S|$, let $S_k = S_{k-1}\setminus\{m_{S_{k-1}}\}$.
3893: (In other words, let $S_k$ consist of the $|S|-k$ largest elements of $S$).
3894: Then $$v_S + \displaystyle{\sum_{k=1}^{|S|-1}} \hs 2^{k-1}\hs v_{S_k}
3895: = \displaystyle{\sum_A}\hs 2^{|\asb|}\hs d_A.$$
3896: \end{claim}
3897:
3898: \begin{proof}
3899: We proceed by induction to show that
3900: $$v_S + \displaystyle{\sum_{k=1}^l} \hs 2^{k-1}\hs v_{S_k} =
3901: \displaystyle{\sum_A} \hs 2^{|\asb|}\hs d_A -
3902: \hs 2^l\cdot\!\displaystyle{\sum_{\overline{S^c_{l+1}}\subs A}}
3903: 2^{|\asb_l|}\hs d_A.$$
3904: The case $l=|S|-1$ is the statement of the claim.
3905: The base case $l=0$ follows from Claim \ref{vs},
3906: together with the observation that $\overline{S^c_{1}}=S^c$.
3907: More generally, for all $l\geq 1$, we have
3908: $$\overline{S^c_{l+1}} = S^c\cup\{m_{S_1},\ldots,m_{S_l}\}.$$
3909: Then
3910: \begin{eqnarray*}
3911: v_S + \displaystyle{\sum_{k=1}^{l+1}} \hs 2^{k-1}\hs v_{S_k} &=&
3912: v_S + \displaystyle{\sum_{k=1}^l}\hs 2^{k-1}\hs v_{S_k} + 2^l\hs v_{S_{l+1}}\\
3913: % &=& \displaystyle{\sum_A} \hs 2^{|\asb|}\hs d_A
3914: % - 2^l\hs\displaystyle{\sum_{\overline{S^c_{l+1}}\subs A}}
3915: % 2^{|\asb_l|}\hs d_A + 2^l\hs v_{S_{l+1}}\hs\hs\hs\hs\text{inductive hypothesis}\\
3916: &=& \displaystyle{\sum_A} \hs 2^{|\asb|}\hs d_A - 2^l\cdot\!\displaystyle{\sum_{\overline{S^c_{l+1}}\subs A}}
3917: 2^{|\asb_l|}\hs d_A + 2^l\cdot\!\sum_{\ss{\overline{S^c_{l+1}}\subs A \\ m_{S_{l+1}}\notin A}}
3918: 2^{|\asb_{l+1}|}\hs d_A
3919: \end{eqnarray*}
3920: by the inductive hypothesis and Claim \ref{vs}.
3921: Using the fact that
3922: $\asb_{l+1}=\asb_l$ when $m_{S_{l+1}}\notin A$,
3923: this is equal to
3924: $$\sum_A\hs 2^{|\asb|}\hs d_A - 2^l\cdot\!\sum_{\overline{S^c_{l+1}}\cup\{m_{S_{l+1}}\}\subs A}
3925: 2^{|\asb_l|}.$$
3926: Finally, since $|\asb_{l+1}|=|\asb_l|-1$ when $m_{S_{l+1}}\in A$,
3927: this reduces to
3928: $$\sum_A \hs 2^{|\asb|}\hs d_A - 2^{l+1}\hs\cdot\!\displaystyle{\sum_{\overline{S^c_{l+2}}
3929: \subs A}}
3930: 2^{|\asb_{l+1}|}\hs d_A,$$ thus proving our claim.
3931: \end{proof}
3932:
3933: For all short subsets $T$ containing $1$,
3934: let $w_T=\displaystyle{\sum_A}2^{|A\cap \bar{T}|}\hs d_A$,
3935: which by Claim \ref{ws} is expressible as a linear combination of elements
3936: of the set $\{v_S\mid \emptyset\neq S\in\mathcal{S}\}$.
3937: Let $$x_S =
3938: \begin{cases}
3939: \displaystyle{\sum_{1\in T\subs S}}(-1)^{|S|+|T|}w_T & \text{if $1\in S$,}\\
3940: v_S & \text{if $1\in S^c$.}
3941: \end{cases}$$
3942: Our last task will be to prove that the transition matrix $\Upsilon$
3943: taking the basis $\{d_A\}$ to
3944: the set $\{x_S\}$
3945: is lower triangular with ones on the diagonal, and therefore invertible.
3946: In order to make sense of ``the diagonal,"
3947: we must first give an explicit bijection between the set of proper subsets of $\twn$
3948: and the set of nonempty short subsets of $\otn$.
3949: We do this as follows: given $A\subsetneq\twn$, let
3950: $$S(A) = \begin{cases}
3951: \Ac &\text{ if $\Ac$ is short,}\\
3952: \otn\setminus\Ac = A\cup\{1\} &\text{ if $\Ac$ is long.}
3953: \end{cases}$$
3954: The rows of $\Upsilon$ will be indexed by $A$,
3955: and the sets will appear in lexicographic order within cardinality
3956: class. For example, when $n=4$, the order of the rows will be
3957: $\emptyset$, $\{2\}$, $\{3\}$, $\{4\}$,
3958: $\{2,3\}$, $\{2,4\}$, $\{3,4\}$.
3959: The columns will be indexed by $S$ according to the bijection described above.
3960:
3961: \begin{claim}\label{tri}
3962: The matrix $\Upsilon$ is lower triangular with ones on the diagonal.
3963: \end{claim}
3964:
3965: \begin{proof}
3966: First consider a column corresponding to a short subset $S$ that does {\it not} contain 1.
3967: The entries in this column correspond to the coefficient of $d_A$ in $x_S = v_S$.
3968: From Claim \ref{vs}, we see that $d_A$ appears in $v_S$ only if $\Lb\subs A\subs \Lb\cup\Sb$,
3969: and if so it appears with a coefficient of $2^{|\asb|}$.
3970: Since $1\notin S$, we have $\Lb = S^c\setminus\{1\} = \twn\setminus S$.
3971: The diagonal entry corresponds to the set $A = \twn\setminus S = \Lb$, therefore
3972: in this row we get the number $2^{|\asb|} = 2^{|\Lb\cap\Sb|} = 1$.
3973: Since the set $A$ corresponding to a given row can never contain the set $B$ corresponding
3974: to a
3975: lower row, the rows above the diagonal fail to satisfy the condition $\Lb\subs A$,
3976: and we get all zeros.
3977:
3978: Now consider a column corresponding to a short subset $S$ that {\it does} contain $1$.
3979: In this case, the coefficient of $d_A$ in $x_S$ is
3980: $$(-1)^{|S|}\displaystyle{\sum_{1\in T\subs S}}(-1)^{|T|}2^{|A\cap\bar{T}|}.$$
3981: The diagonal entry corresponds to the set $A=\Sb$, and we get
3982: \begin{eqnarray*}
3983: (-1)^{|S|}\displaystyle{\sum_{1\in T\subs S}}(-1)^{|T|}2^{|\bar{T}|}
3984: &=& (-1)^{|\Sb|}\displaystyle{\sum_{1\in T\subs S}}(-2)^{|\bar{T}|}\\
3985: &=& (-1)^{|\Sb|}(1-2)^{|\Sb|} = 1.
3986: \end{eqnarray*}
3987: Any row above the diagonal corresponds to a set $A$ which does not contain $\Sb$.
3988: Choose an element $l\in\Sb\setminus A$. Then
3989: \begin{eqnarray*}
3990: (-1)^{|S|}\displaystyle{\sum_{1\in T\subs S}}(-1)^{|T|}2^{|A\cap\bar{T}|}
3991: &=& (-1)^{|S|}\displaystyle{\sum_{l\in T}}(-1)^{|T|}2^{|A\cap\bar{T}|}
3992: + (-1)^{|S|}\displaystyle{\sum_{l\notin T}}(-1)^{|T|}2^{|A\cap\bar{T}|}\\
3993: &=& (-1)^{|S|}\displaystyle{\sum_{l\notin T}}\left[(-1)^{|T|}2^{|A\cap\bar{T}|}
3994: + (-1)^{|T\cup\{l\}|}2^{|A\cap\bar{T}|}\right]\\
3995: &=& 0.
3996: \end{eqnarray*}
3997: Hence $\Upsilon$ is lower triangular.
3998: \end{proof}
3999:
4000: Claim \ref{tri} tells us that each $d_A$ can be expressed as a linear
4001: combination of elements of the form $x_S$, and therefore of elements
4002: of the form $v_S$. This completes the proof of Lemma \ref{tech}
4003: \end{prooftech}
4004: \end{section}
4005:
4006: \begin{section}{Cohomology of the core components}\label{last}
4007: In this section we compute the $S^1$-equivariant
4008: and ordinary cohomology rings of the core component
4009: $U_S$ corresponding to a short subset $S\subs\otn$. Since $U_S$ is the closure of a cell
4010: in an even-dimensional equivariant cellular decomposition of $\M$, the restriction
4011: map $\hso(\M)\to\hso(U_S)$ is surjective. In particular, $\hso(U_S)$
4012: is generated by restrictions
4013: of the Kirwan classes $c_1,\ldots,c_n,x$.
4014: For our presentation, it will be convenient
4015: to assume that $1\in S$, and to work with the classes
4016: $d_k = \half(c_1+c_k)$ introduced in the proof of Lemma \ref{tech}.
4017: With respect to these generators, we obtain the following result.
4018:
4019: \begin{theorem}\label{eqcore}
4020: The equivariant cohomology ring $\hso(U_S)$ is isomorphic to
4021: $\Q[d_1,\ldots,d_n,x]/\mathcal{J}_S,$
4022: where $\mathcal{J}_S$ is generated by the following four families:
4023: \begin{eqnarray*}&1)&\hs\hs d_1 - d_i \hs\hs\text{ for all }\hs\hs i\in S\\
4024: &2)&\hs\hs
4025: d_j(d_1 - d_j)
4026: \hs\hs\text{ for all }\hs\hs j\in S^c\\
4027: &3)&\hs\hs \prod_{j\in R}d_j\hs\hs\text{ for all }\hs\hs R\subs S^c\text{ such that }R\cup S
4028: \text{ is long}\\
4029: &4)&\hs\hs (d_1 + x)^{|S|-1}\cdot\frac{1}{d_1}\left(\prod_{j\in L}(d_j-d_1)-\prod_{j\in L}d_j\right)
4030: \hs\hs\text{ for all long subsets }L\subs S^c.
4031: \end{eqnarray*}
4032: \end{theorem}
4033:
4034: \begin{corollary}\label{ordcore}
4035: The ordinary cohomology ring $H^*(U_S)$ is isomorphic to
4036: $\Q[d_1,\ldots,d_n]/\mathcal{I}_S,$
4037: where $\mathcal{I}_S$ is generated by the following four families:
4038: \begin{eqnarray*}&1)&\hs\hs d_1 - d_i \hs\hs\text{ for all }\hs\hs i\in S\\
4039: &2)&\hs\hs
4040: d_j(d_1 - d_j)
4041: \hs\hs\text{ for all }\hs\hs j\in S^c\\
4042: &3)&\hs\hs \prod_{j\in R}d_j\hs\hs\text{ for all }\hs\hs R\subs S^c\text{ such that }R\cup S
4043: \text{ is long}\\
4044: &4)&\hs\hs d_1^{|S|-2}\prod_{j\in L}(d_j-d_1)
4045: \hs\hs\text{ for all long subsets }L\subs S^c.
4046: \end{eqnarray*}
4047: \end{corollary}
4048:
4049: \begin{remark}\label{integer}
4050: Each of these relations has a geometric interpretation.
4051: For $i\in\{1,\ldots,n\}$, it is possible to construct a line bundle on $\M$ with equivariant
4052: Euler class $d_i - d_1$
4053: which has a section supported on the locus
4054: where $q_1q_1^*$ and $q_iq_i^*\in\R^3$ point in opposite directions.
4055: Since this locus is disjoint
4056: from $U_S$ when $i\in S$, we have $$1)\hs\hs d_i = d_1 \in \hso(\us)\hs\text{ for all }\hs i\in S.$$
4057: Similarly,
4058: $-d_j = -\half(c_1+c_j)$ is represented by the divisor $Z_{1j}\subs\M$ of points
4059: on which $q_1q_1^*$ and $q_iq_i^*\in\R^3$ point in the same direction
4060: \cite[\S 3]{HP2}.
4061: Then by the previous reasoning, we obtain
4062: $$2)\hs\hs d_j(d_1 - d_j) = 0\in\hso(\us)\hs\text{ for all }\hs j\in S^c.$$
4063: For any $R\subs S^c$, we may intersect the divisors $Z_{ij}\subs\M$
4064: (defined in the analogous way) for all $j\in R$
4065: to find that
4066: the cohomology class $(-1)^{|R|}\prod_{j\in R}d_j$ is represented
4067: by the subvariety $Z_R\subs \M$ of points with $q_j$ proportional to $q_1$
4068: for all $j\in R$.
4069: When restricted to $U_S$, this becomes $U_S\cap U_{R\cup S}$,
4070: the closure of the unstable manifold for the critical locus $\M_{R\cup S}\cap U_S$
4071: of the Morse-Bott function $\Phi|_{U_S}$.
4072: In particular, we have $$3)\hs\hs \prod_{j\in R}d_j=0\in\hso(\us)\text{ if }R\cup S\hs\text{ is long}.$$
4073: To understand the fourth family of relations, we note that the class
4074: $$d_1+x = 2d_i-d_1+x = c_i+x \in \hso(U_S)$$
4075: is represented by the divisor $W_i$ of points
4076: with $p_i = 0$ for any $i\in S$ \cite[\S 3]{HP2}. In particular,
4077: $(d_1+x)^{|S|-1}$
4078: is represented by the subvariety of points in $U_S$ on which $p_i = 0$
4079: for all $i\in\Sb$, which is equal to $\X_S$ by the complex moment map condition.
4080: Hence the fourth family of generators of $\mathcal{J}_S$ (or of $\mathcal{I}_S$)
4081: can be interpreted geometrically as $(d_1+x)^{|S|-1}$
4082: (respectively $d_1^{|S|-1}$
4083: in the nonequivariant case) times
4084: classes that vanish in $\hso(\X_S)$ (see Lemma \ref{pols}).
4085: %
4086: % The critical submanifold $J_{R\cup S}\cap U_S$ is a projective
4087: % space with cohomology ring generated over $\Z[x]$ by the restriction
4088: % of $d_1+x$. It is shown in $\cite{HK}$ that the cohomology ring
4089: % of $\X_S$ is also generated over the integers by the restrictions
4090: % of $d_1,\ldots,d_n$. Hence Morse-Bott theory tells us that
4091: % $d_1,\ldots,d_n,x$ is actually an {\em integer} basis for $\hso(U_S)$.
4092: \end{remark}
4093:
4094: \begin{eqcoreproof}
4095: Let $\phi:\Q[d_1,\ldots,d_n,x]\to\hso(\us)$ denote the composition of the Kirwan
4096: map with restriction to $\us$. Our claim is that $\ker\phi = \mathcal{J}_S$.
4097: For every short subset $T$ containing $S$,
4098: let $$\phi_T:\Q[d_1,\ldots,d_n,x]\to\hso(\mtus)$$ denote
4099: the composition of the Kirwan map with restriction to $\mtus$,
4100: and let $$J_T = \ker\phi_T.$$
4101: Similarly, let $$\phi_{\emptyset}:\Q[d_1,\ldots,d_n,x]\to\hso(\X_S)$$
4102: be the natural map, and let
4103: $$J_{\emptyset} = \ker\phi_{\emptyset}.$$
4104: The kernel of the restriction map $\hso(U_S)\to\hso(U_S^{S^1})$
4105: to the fixed point set of $U_S$ is a torsion module over $\hso(pt)$ \cite[3.5]{AB},
4106: and Proposition \ref{formality}
4107: tells us that $\hso(U_S)$ is a free $\hso(pt)$-module.
4108: Hence the restriction map is injective,
4109: and we have $$\ker\phi = \ker\phi_{\emptyset}\cap\bigcap_{\ts}\ker\phi_T.$$
4110: We know that $\mtus\cong\C P^{|S|-2}$ for all short $\ts$,
4111: therefore $$\hso(\mtus)\cong\Q[h,x]/h^{\sz}.$$
4112: Furthermore, we know that for all $i\in T$, the restriction of $d_i + x$
4113: to $\hso(U_T)$ is represented by the divisor $W_i\cap U_T$ (see Remark \ref{integer}),
4114: and therefore restricts to the class of a hyperplane on $\mtus$.
4115: Hence $\phi_T(d_i+x)=h$ for all $i\in T$.
4116: On the other hand, for $j\in T^c$, the class $d_j$
4117: is represented by the divisor $Z_{1j}$ on $\M$, which is disjoint from $\mtus$,
4118: hence $\phi(d_j) = 0$ for all $j\in T^c$.
4119: Thus we conclude that
4120: $$\ker\phi_T = \<d_1-d_i, \hs d_j, (d_1+x)^{\sz}\mid i\in T, j\in T^c\>.$$
4121:
4122: \begin{lemma}\label{jt}
4123: The intersection
4124: $\displaystyle{
4125: \bigcap_{\ts}}\ker\phi_T$
4126: is equal to
4127: $$\<d_1-d_i, \hs d_j(d_1 - d_j),
4128: \displaystyle{\prod_{j\in R}d_j},
4129: (d_1+x)^{\sz}
4130: % where $i$ ranges over $S$, $j$ ranges over $S^c$,
4131: % and $R\subs S^c$ is any subset such that $R\cup S$ is long.
4132: \hs\hs\bigg |\hs\hs i\in S, j\in S^c, R\cup S\text{ long}\>.$$
4133: \end{lemma}
4134:
4135: \begin{proof}
4136: First, since the variable $x$ appears only in the generator $(d_1+x)^{\sz}$,
4137: which is contained in every ideal on both sides of the equation, we may reduce
4138: the problem to showing that
4139: \begin{equation}\label{reduce}
4140: \bigcap_{\ts}\big\langle d_1-d_i, \hs d_j\bigmid i\in T, j\in T^c\big\rangle =
4141: \<d_1-d_i, \hs d_j(d_1-d_j), \prod_{j\in R}d_j
4142: \hs\hs\bigg |\hs\hs i\in S, j\in S^c, R\cup S\text{ long}\>
4143: \end{equation}
4144: in the ring $\Q[d_1,\ldots,d_n]$.
4145: Both ideals cut out the (reducible) variety
4146: $$\bigcup_{\ts}Y_T\subs\operatorname{Spec}\Q[d_1,\ldots,d_n],$$
4147: where $$Y_T = \big\{(z_1,\ldots,z_n\bigmid
4148: z_i=z_1\hs\forall i\in S, z_j=0\hs\forall j\in S^c\big\}.$$
4149: The left hand side of Equation \eqref{reduce} is an intersection of prime ideals,
4150: and is therefore radical.
4151: Thus by Hilbert's Nullstellensatz, it is sufficient to prove that
4152: the right hand side of Equation \eqref{reduce} is radical.
4153: This involves showing that the ideal is saturated, with
4154: Hilbert polynomial equal to the constant $\#\{\text{short }\ts\}$.
4155:
4156: The degree $k$ piece of the quotient
4157: $$\Q[d_1,\ldots,d_n]/\<d_1-d_i, \hs d_j(d_1-d_j)
4158: \mid i\in S, j\in S^c\>$$
4159: has a basis of elements of the form
4160: $$d_1^{e_1}\prod_{j\in S^c}d_j^{e_j},$$
4161: where $e_j \in\{0,1\}$ for all $j>0$, and $e_1+\sum_{j\in S^c}e_j = k$.
4162: The subset of these elements with the property that
4163: $S\cup\{j\mid e_j = 1\}$ is short descends to a basis for the
4164: degree $k$ part of the ring
4165: $$\Q[d_1,\ldots,d_n]\bigg/\<d_1-d_i, \hs d_j(d_1-d_j), \prod_{j\in R}d_j
4166: \hs\hs\bigg |\hs\hs i\in S, j\in S^c, R\cup S\text{ long}\>,$$
4167: hence our ideal has the desired Hilbert polynomial.
4168: It is also clear from this description that if an element
4169: $a$ of the quotient ring is nonzero, so is $d_1^d\cdot a$ for any $d\geq 0$,
4170: hence our ideal is saturated.
4171: \end{proof}
4172:
4173: It now remains to show that
4174: $$\mathcal{J}_S = \<d_1-d_i, \hs d_j(d_1 - d_j), \prod_{j\in R}d_j,
4175: \hs (d_1+x)^{\sz} \hs\hs\bigg |\hs\hs i\in S, j\in S^c, R\cup S\text{ long}\> \cap \ker\phi_{\emptyset}.$$
4176: The fact that $\mathcal{J}_S$ is contained in the intersection is clear.
4177: To show the opposite containment, consider an element
4178: $$a + \eta\cdot (d_1+x)^{\sz}\in \<d_1-d_i, \hs d_j(d_1 - d_j), \prod_{j\in R}d_j,
4179: \hs (d_1+x)^{\sz} \hs\hs\bigg |\hs\hs i\in S, j\in S^c, R\cup S\text{ long}\>,$$
4180: with $$a\in \<d_1-d_i, \hs d_j(d_1 - d_j), \prod_{j\in R}d_j\hs\hs\bigg |\hs\hs
4181: i\in S, j\in S^c, R\cup S\text{ long}\>,$$
4182: and suppose that we also have
4183: $$a + \eta\cdot (d_1+x)^{\sz}\in\ker\phi_{\emptyset}.$$
4184:
4185: \begin{lemma}\label{pols}{\em \cite{HK}}
4186: The kernel of $\phi_{\emptyset}$
4187: is equal to
4188: $$\<d_1-d_i, \hs d_j(d_1-d_j), \prod_{j\in R}d_j,
4189: \hs (d_1+x)^{\sz}d_1^{-1}\left(\prod_{j\in L}(d_j-d_1)-\prod_{j\in L}d_j\right)\>,$$
4190: where $i\in S, j\in S^c$, and $R,L\subs S^c$, with $R\cup S$ and $L$ both long.
4191: % \mid i\in S, j\in S^c, R\cup S\text{ long}, L\subs S^c\text{ long}\>.$$
4192: \end{lemma}
4193:
4194: Lemma \ref{pols} tells us that $a\in \ker\phi_{\emptyset}$,
4195: therefore $$\eta\cdot (d_1+x)^{\sz}\in\ker\phi_{\emptyset}.$$
4196: But $(d_1+x)^{\sz}$ is represented in $\hso(U_S)$ by the subvariety
4197: $\X_S$ (see Remark \ref{integer}), hence
4198: $$0 = \phi_{\emptyset}(\eta\cdot (d_1+x)^{\sz})
4199: = \phi_{\emptyset}(\eta)\cdot e(\X_S),$$
4200: where $e(\X_S)$ is the equivariant Euler class of the normal bundle
4201: to $\X_S$ inside of $U_S$. Since the equivariant Euler class
4202: of the normal bundle to a component of the fixed point set is never a zero-divisor,
4203: we have $\eta\in\ker\phi_{\emptyset}$.
4204: Then by Equation \ref{pols}, $$a+\eta\cdot (d_1+x)^{\sz}\in\mathcal{J}_S.$$
4205: This completes the proof of Theorem \ref{eqcore}.
4206: \end{eqcoreproof}
4207:
4208: \vspace{-\baselineskip}
4209: \begin{example}\label{projective}
4210: For arbitrary $n$ and $\a$, suppose that $S$ is a maximal short subset.
4211: Then Corollary \ref{ordcore} tells us that
4212: $H^*(U_S)\cong\Q[d_1]/\langle d_1^{n-2}\rangle$.
4213: We conjecture that in this case we in fact have $U_S\cong \C P^{n-3}$.
4214: \end{example}
4215:
4216: \begin{example}\label{again}
4217: Consider the core component pictured in Example \ref{ooh}.
4218: By Theorem \ref{ordcore} and Remark \ref{integer},
4219: $$H^*(U_S) \cong \Q[d_1,d_3,d_4,d_5]
4220: \Bigg/
4221: \left<\begin{array}{c}
4222: d_3(d_1-d_3),\hs d_4(d_1-d_4),\hs d_5(d_1-d_5),\hs d_3d_4,\hs d_3d_5,\hs d_4d_5,\\
4223: d_1(d_1-d_3-d_4),\hs d_1(d_1-d_3-d_5),\hs d_1(d_1-d_4-d_5)
4224: \end{array}\right>,$$
4225: where $d_1$ is the fundamental class of $\X_S$,
4226: and $d_3$, $d_4$, and $d_5$ are the negatives of the fundamental classes of the curves
4227: labeled $123$, $124$, and $125$, respectively.
4228: Because the transverse intersection of two complex varieties is positive,
4229: we know that $-d_1d_3[U_S] = 1$.
4230: With respect to the basis $$\{d_1-d_3-d_4-d_5,\, d_3,\, d_4,\, d_5\},$$
4231: the intersection form on $H^2(U_S)$ is represented by the matrix
4232: $$\left(\begin{array}{cccc}
4233: 1&&&\\
4234: &-1&&\\
4235: &&-1&\\
4236: &&&-1\end{array}\right).$$
4237: It is likely that $U_S$ is isomorphic
4238: to the blow-up of $\C P^2$ at three points.
4239: \end{example}
4240:
4241: \begin{example}
4242: Using the same $\a= (1,1,3,3,3)$,
4243: consider the short subset $S=\{1,3\}$.
4244: In this case, Theorem \ref{ordcore} tells us that
4245: $$H^*(U_S)\cong\Q[d_1,d_2]/\left<d_1^2, d_2(d_1-d_2)\right>.$$
4246: With respect to the basis $\{d_1-d_2,\, d_2\}$,
4247: the intersection form on $H^2(U_S)$ is represented by the matrix
4248: $\footnotesize{\Big(\begin{array}{cc}
4249: -1&0\\
4250: 0& 1
4251: \end{array}
4252: \Big)}$,
4253: hence $U_S$ is homeomorphic to the blow-up of $\C P^2$ at a single point.
4254: \end{example}
4255: \end{section}
4256: \end{chapter}
4257:
4258: \footnotesize{
4259: \begin{thebibliography}{10}
4260:
4261: \bibitem[AB]{AB}
4262: M.~Atiyah and R.~Bott.
4263: \newblock The moment map and equivariant cohomology.
4264: \newblock {\em Topology} 23 (1984), 1--28.
4265:
4266: \bibitem[BV]{BV} N.~Berline and M.~Vergne.
4267: \newblock Z\'eros d'un champ de vecteurs et classes caract\'eristiques
4268: \'equivariantes.
4269: \newblock {\em Duke Math. J.} 50 (1983) no. 2, 539--549.
4270:
4271: \bibitem[Bi]{Bi}
4272: R.~Bielawski.
4273: \newblock Complete hyperkaehler $4n$-manifolds with a local tri-Hamiltonian $R^n$-action.
4274: \newblock {\em Math. Ann.} 314 (1999) no. 3, 505--528.
4275:
4276: \bibitem[BD]{BD}
4277: R.~Bielawski and A.~Dancer.
4278: \newblock The geometry and topology of toric hyperk\"ahler manifolds.
4279: \newblock {\em Comm. Anal. Geom.} 8 (2000), 727--760.
4280:
4281: \bibitem[BGH]{BGH}
4282: D.~Biss, T.~Holm, and V.~Guillemin.
4283: \newblock The mod 2 cohomology of fixed point sets
4284: of anti-symplectic involutions.
4285: \newblock ArXiv: math.SG/0107151.
4286:
4287: \bibitem[BJ]{BJ}
4288: T.~Brocker and K.~Janich.
4289: \newblock {\em Introduction to Differential Topology}.
4290: \newblock Cambridge University Press, 1982.
4291:
4292: \bibitem[De]{De}
4293: T.~Delzant.
4294: \newblock Hamiltoniens p\'eriodiques et images convexes de l'application moment.
4295: \newblock {\em Bull. Soc. Math. France} 116 (1988), no. 3, 315--339.
4296:
4297: \bibitem[Do]{Do}
4298: I.~Dolgachev.
4299: \newblock {\em Lectures on invariant theory}.
4300: \newblock London Mathematical Society Lecture Note Series, 296.
4301: Cambridge University Press, 2003.
4302:
4303: \bibitem[ES]{ES}
4304: G.~Ellingsrud and S.A.~Str\o mme.
4305: \newblock On the Chow ring of a geometric quotient.
4306: \newblock {\em Ann. of Math.} (2) 130 (1989), no. 1, 159--187.
4307:
4308: \bibitem[Fu]{Fu}
4309: W.~Fulton.
4310: \newblock Eigenvalues, invariant factors, highest weights,
4311: and Schubert calculus.
4312: \newblock {\em Bull. Amer. Math. Soc.} 37 (2000), no. 3, 209--249 .
4313:
4314: \bibitem[M2]{M2}
4315: D.~Grayson and M.~Stillman.
4316: \newblock Macaulay 2: a computer algebra system for algebraic geometry.
4317: \newblock Version 0.8.56, www.math.uiuc.edu/macaulay2, 1999.
4318:
4319: \bibitem[GS]{GS}
4320: V.~Guillemin and S.~Sternberg.
4321: \newblock The coefficients of the Duistermaat-Heckman polynomial
4322: and the cohomology ring of reduced spaces.
4323: \newblock {\em Geometry, topology, \& physics}, 202--213,
4324: Conf. Proc. Lecture Notes Geom. Topology, IV,
4325: Internat. Press, Cambridge, MA, 1995.
4326:
4327: \bibitem[HH]{HH}
4328: M.~Harada and T.~Holm.
4329: \newblock The equivariant cohomology
4330: of hypertoric varieties and their real loci.
4331: \newblock In preparation.
4332:
4333: \bibitem[HP1]{HP1}
4334: M.~Harada and N.~Proudfoot.
4335: \newblock Properties of the residual circle action on a hypertoric
4336: variety.
4337: \newblock To appear in {\em Pacific J. Math.}
4338: \newblock ArXiv: math.DG/0207012.
4339:
4340: \bibitem[HP2]{HP2}
4341: M.~Harada and N.~Proudfoot.
4342: \newblock Hyperpolygon spaces and their cores.
4343: \newblock To appear in {\em Trans. A.M.S.}
4344: \newblock ArXiv: math.AG/0308218.
4345:
4346: \bibitem[H1]{H1}
4347: T.~Hausel.
4348: \newblock Geometry of the moduli space of Higgs bundles.
4349: \newblock Thesis for Ph.D. in Pure Mathematics,
4350: DPMMS, Cambridge University, August 1998.
4351: \newblock ArXiv: math.AG/0107040.
4352:
4353: \bibitem[H2]{H2}
4354: T.~Hausel.
4355: \newblock Compactification of moduli of Higgs bundles.
4356: \newblock {\em J. Reine Angew. Math.} 503 (1998), 169--192.
4357: \newblock ArXiv: math.AG/9804083.
4358:
4359: \bibitem[H3]{H3}
4360: T.~Hausel.
4361: \newblock Quaternionic geometry of matroids.
4362: \newblock ArXiv: math.AG/0308146.
4363:
4364: \bibitem[HP]{HP}
4365: T.~Hausel and N.~Proudfoot.
4366: \newblock Abelianization for hyperk\"ahler quotients.
4367: \newblock ArXiv: math.SG/0310141.
4368:
4369: \bibitem[HS]{HS}
4370: T.~Hausel and B.~Sturmfels.
4371: \newblock Toric hyperk\"ahler varieties.
4372: \newblock {\em Doc. Math.} 7 (2002), 495--534 (electronic).
4373: \newblock ArXiv: math.AG/0203096.
4374:
4375: \bibitem[HT1]{HT1} T.~Hausel and M.~Thaddeus.
4376: \newblock Generators for the cohomology ring of the
4377: moduli space of rank 2 Higgs bundles.
4378: \newblock To appear in {\em Proc. London Mat. Soc.}
4379: \newblock ArXiv: math.AG/0003093.
4380:
4381: \bibitem[HT2]{HT2} T.~Hausel and M.~Thaddeus.
4382: \newblock Relations in the cohomology ring of the
4383: moduli space of rank 2 Higgs bundles,
4384: \newblock {\em J. Amer. Math. Soc.}, 16 (2003), 303--329
4385: \newblock ArXiv: math.AG/0003094.
4386:
4387: \bibitem[HK]{HK}
4388: J-C.~Hausmann, A.~Knutson.
4389: \newblock The cohomology ring of polygon spaces.
4390: \newblock {\em Ann. Inst. Fourier, Grenoble} 48 (1998) no. 1, 281--321.
4391:
4392: % \bibitem[HL]{HL}
4393: % P.~Heinzner and F.~Loose.
4394: % \newblock Reduction of complex Hamiltonian $G$-spaces.
4395: % \newblock {\em Geom. Funct. Anal.} 4 (1994), no. 3, 288--297.
4396: %
4397: \bibitem[Hi]{Hi}
4398: N.~Hitchin.
4399: \newblock The self-duality equations on a Riemann surface.
4400: \newblock {\em Proc. London Math. Soc.} (3) 55 (1987) no. 1, 59--126.
4401:
4402: \bibitem[HKLR]{HKLR}
4403: N.~Hitchin, A.~Karlhede, U.~Lindstr\"om, and M.~Ro\v{c}ek.
4404: \newblock Hyper-K\"ahler metrics and supersymmetry.
4405: \newblock {\em Comm. Math. Phys.} 108 (1987)
4406: no. 4, 535--589.
4407:
4408: \bibitem[KP]{KP}
4409: A.~Khovanskii and A.~Pukhlikov.
4410: \newblock The Riemann-Roch theorem for integrals and sums of quasipolynomials
4411: on virtual polytopes.
4412: \newblock {\em St. Petersburg Math. J.} 4 (1993), 789--812.
4413:
4414: \bibitem[Ki]{Ki}
4415: F.~Kirwan.
4416: \newblock {\em Cohomology of quotients in symplectic and algebraic geometry}.
4417: \newblock Mathematical Notes, 31.
4418: \newblock Princeton University Press, 1984.
4419:
4420: \bibitem[Kl]{Kl}
4421: A.~Klyachko.
4422: \newblock Spatial polygons and stable configurations of points in the projective line.
4423: \newblock Algebraic geometry and its applications (Yaroslavl, 1992),
4424: {\em Aspects Math.}, Vieweg, Braunschweig (1994) 67--84.
4425:
4426: \bibitem[KT]{KT}
4427: A.~Knutson and T.~Tao.
4428: \newblock
4429: The honeycomb model of $GL(n)$ tensor products I:
4430: proof of the saturation conjecture.
4431: \newblock {\em Journal of the A.M.S.} {\bf 12} (1999) no. 4, 1055--1090.
4432:
4433: \bibitem[K1]{K1}
4434: H.~Konno.
4435: \newblock Cohomology rings of toric hyperk\"ahler manifolds.
4436: \newblock {\em Int. J. of Math.} 11 (2000)
4437: no. 8, 1001--1026.
4438:
4439: \bibitem[K2]{K2}
4440: H.~Konno.
4441: \newblock On the cohomology ring of the HyperK\"ahler analogue of the
4442: Polygon Spaces.
4443: \newblock Integrable systems, topology, and physics (Tokyo, 2000), 129--149,
4444: {\em Contemp. Math.}, 309,
4445: Amer. Math. Soc., Providence, RI, 2002.
4446:
4447: \bibitem[Le]{Le}
4448: E.~Lerman.
4449: \newblock Symplectic cuts.
4450: \newblock {\em Math. Res. Lett.} 2 (1995) no. 3, 247--258.
4451:
4452: \bibitem[LT]{LT}
4453: E.~Lerman and S.~Tolman.
4454: \newblock Hamiltonian torus actions on symplectic orbifolds and toric varieties.
4455: \newblock {\em Trans. Amer. Math. Soc.} 349 (1997) no. 10, 4201--4230.
4456:
4457: \bibitem[Mk]{Mk}
4458: E.~Markman.
4459: \newblock Generators of the cohomology ring of moduli spaces
4460: of sheaves on symplectic surfaces.
4461: \newblock {\emph J. Reine Angew. Math.} 544 (2002), 61--82.
4462:
4463: \bibitem[Ma]{Ma}
4464: S.~Martin.
4465: \newblock Symplectic quotients by a nonabelian group and
4466: by its maximal torus.
4467: \newblock To appear in the {\em Ann. of Math.}
4468: \newblock ArXiv: math.SG/0001002.
4469:
4470: \bibitem[MNS]{MNS}
4471: G.~Moore, N.~Nekrasov, and S.~Shatashvili.
4472: \newblock \emph{Integrating over Higgs branches}.
4473: \newblock {\em Comm. Math. Phys.} 209 (2000) no.~1, 97--121.
4474:
4475: \bibitem[MFK]{MFK}
4476: D.~Mumford, J.~Fogarty, and F.~Kirwan.
4477: \newblock {\em Geometric Invariant Theory}.
4478: \newblock Third edition. Ergebnisse der Mathematik und ihrer
4479: Grenzgebiete (2) [Results in Mathematics and Related Areas (2)], 34.
4480: \newblock Springer-Verlag, 1994.
4481:
4482: \bibitem[N1]{N1}
4483: H.~Nakajima.
4484: \newblock Instantons on ALE spaces, quiver varieties, and Kac-Moody algebras.
4485: \newblock {\em Duke Math. J.} 76 (1994) no. 2, 365--416.
4486:
4487: \bibitem[N2]{N2}
4488: H.~Nakajima.
4489: \newblock Varieties associated with quivers.
4490: \newblock {\em Canadian Math. Soc. Conf. Proc.} 19 (1996), 139--157.
4491:
4492: \bibitem[N3]{N3}
4493: H.~Nakajima.
4494: \newblock Quiver varieties and Kac-Moody algebras.
4495: \newblock {\em Duke Math. J.} 91 (1998) no. 3, 515--560.
4496:
4497: \bibitem[N4]{N4}
4498: H.~Nakajima.
4499: \newblock {\em Lectures on Hilbert schemes of points on surfaces}.
4500: \newblock University Lecture Series, 18.
4501: \newblock American Mathematical Society, 1999.
4502:
4503: \bibitem[N5]{N5}
4504: H.~Nakajima.
4505: \newblock Quiver varieties and finite-dimensional representations of quantum affine algebras.
4506: \newblock {\em J. Amer. Math. Soc.} 14 (2001) no. 1, 145--238.
4507:
4508: \bibitem[NS]{NS}
4509: M.~Narasimhan and C.~Seshadri.
4510: \newblock Stable bundles and unitary bundles on a compact Riemann surface.
4511: \newblock {\em Proc. Nat. Acad. Sci. U.S.A.} 52 (1964) 207--211.
4512:
4513: \bibitem[OS]{OS}
4514: P.~Orlik and L.~Solomon.
4515: \newblock Combinatorics and topology of complements of hyperplanes.
4516: \newblock {\em Invent. Math.} 56 (1980), 167--189.
4517:
4518: \bibitem[OT]{OT}
4519: P.~Orlik and H.~Terao.
4520: \newblock {\em Arrangements of Hyperplanes}.
4521: \newblock Grundlehren der mathematischen Wissenschaften, 300.
4522: \newblock Springer-Verlag, 1992.
4523:
4524: % \bibitem[Pa]{Pa} P.~Paradan.
4525: % \newblock The moment map and equivariant cohomology with generalized coefficients.
4526: % \newblock\emph{Topology} {39} (2000), no. 2, 401--444.
4527: %
4528: \bibitem[Pr]{Pr}
4529: N.~Proudfoot.
4530: \newblock {\em The equivariant Orlik-Solomon algebra}.
4531: \newblock ArXiv: math.CO/0306013.
4532:
4533: \bibitem[Sc]{Sc}
4534: C.~Schmid.
4535: \newblock {\em Cohomologie \'equivariante des certaines vari\'et\'es
4536: hamiltoniennes et de leur partie r\'eelle.}
4537: \newblock Th\`ese \`a Universit\'e de Gen\`eve.
4538: \newblock Available at http://www.unige.ch/math/biblio/these/theses.html.
4539:
4540: \bibitem[Si]{Si}
4541: C.~Simpson.
4542: \newblock Higgs bundles and local systems.
4543: \newblock {\em Inst. Hautes Études Sci. Publ. Math.} 75 (1992), 5--95.
4544:
4545: % \bibitem[Sj]{Sj}
4546: % R.~Sjamaar.
4547: % \newblock Holomorphic slices, symplecitic reduction, and multiplicities of representations.
4548: % \newblock {\em Ann. Math.} 141 (1995), 87--129.
4549: %
4550: \bibitem[St]{St}
4551: R.~Stanley.
4552: \newblock The number of faces of a simplicial convex polytope.
4553: \newblock {\em Adv. in Math.} 35 (1980), no. 3, 236--238.
4554:
4555: \bibitem[Th]{Th} M.~Thaddeus.
4556: \newblock Private communication.
4557:
4558: \bibitem[TW]{TW}
4559: S.~Tolman and J.~Weitsman.
4560: \newblock Cohomology rings of symplectic quotients.
4561: \newblock {\em Comm. An. Geom.} 11 (2003) no. 4, 751--773.
4562: \end{thebibliography}
4563: }
4564: % \end{spacing}
4565: \end{document}
4566: