math0207012/ht.tex
1: \documentclass[12pt]{article}
2: \usepackage{amsmath}
3: \usepackage{amssymb}
4: \usepackage{amscd}
5: % \usepackage{euler}
6: \usepackage{latexsym}
7: \usepackage{theorem}
8: \usepackage{doublespace}
9: \usepackage{epsfig}
10: \usepackage{psfrag}
11: \usepackage{amsfonts}
12: \usepackage{enumerate}
13: 
14: \newtheorem{theorem}{Theorem}[section]
15: \newtheorem{lemma}[theorem]{Lemma}
16: \newtheorem{proposition}[theorem]{Proposition}
17: \newtheorem{corollary}[theorem]{Corollary}
18: \newtheorem{claim}[theorem]{Claim}
19: \newtheorem{conjecture}[theorem]{Conjecture}
20: 
21: {\theorembodyfont{\rmfamily}
22: \theoremstyle{plain}
23: \newtheorem{definition}[theorem]{Definition}
24: \newtheorem{example}[theorem]{Example}
25: \newtheorem{problem}[theorem]{Problem}
26: \newtheorem{remark}[theorem]{Remark}
27: }
28: 
29: % Margin stuff from jason
30: \oddsidemargin=0pt
31: \evensidemargin=0pt
32: \topmargin=0in
33: \headheight=0pt
34: \headsep=0pt
35: \setlength{\textheight}{9in}
36: \setlength{\textwidth}{6.5in}
37: 
38: % \renewcommand{\familydefault}{ppl}
39: 
40: \newcommand{\bd}{\bf \boldmath}
41: \newcommand{\HH}{{\mathbb H}}
42: \newcommand{\T}{{\mathbb T}}
43: \newcommand{\gl}{{\mathfrak gl}}
44: \newcommand{\symp}{{\mathfrak sp}}
45: \newcommand{\liet}{{\mathfrak t}}
46: \newcommand{\lieu}{{\mathfrak u}}
47: \newcommand{\V}{{\mathbb V}}
48: \newcommand{\F}{{\mathbb F}}
49: \newcommand{\e}{\varepsilon}
50: \newcommand{\Id}{\bf 1}
51: \newcommand{\PP}{{\mathbb P}}
52: 
53: \renewcommand{\k}{\kappa}
54: \newcommand{\ktd}{\kappa_{\Td}}
55: \newcommand{\ktdso}{\kappa_{\Td\times S^1}}
56: \newcommand{\kso}{\kappa_{S^1}}
57: 
58: \newcommand{\<}{\left<}
59: \renewcommand{\>}{\right>}
60: 
61: \renewcommand{\Re}{\operatorname{Re}}
62: \renewcommand{\Im}{\operatorname{Im}}
63: \newcommand{\im}{\operatorname{im}}
64: \newcommand{\id}{\operatorname{id}}
65: \newcommand{\Coker}{\operatorname{Coker}}
66: \newcommand{\Ker}{\operatorname{Ker}}
67: \renewcommand{\dim}{\operatorname{dim}}
68: 
69: \newcommand{\Ot}{\displaystyle\bigotimes}
70: \newcommand{\Od}{\displaystyle\bigoplus}
71: \newcommand{\Pd}{\displaystyle\prod}
72: \newcommand{\Sd}{\displaystyle\sum}
73: \newcommand{\Cd}{\displaystyle\bigcap}
74: 
75: \renewcommand{\H}{{\mathbb{H}}}
76: \newcommand{\C}{{\mathbb{C}}}
77: \newcommand{\Zt}{{\mathbb{Z}_2}}
78: \newcommand{\Z}{{\mathbb{Z}}}
79: \newcommand{\Q}{{\mathbb{Q}}}
80: \newcommand{\R}{{\mathbb{R}}}
81: 
82: \newcommand{\la}{\lambda}
83: \renewcommand{\a}{\alpha}
84: \renewcommand{\b}{\beta}
85: \renewcommand{\t}{\theta}
86: \newcommand{\s}{\sigma}
87: 
88: \newcommand{\otd}{\{1,\ldots,d\}}
89: \newcommand{\Hn}{{\H^n}}
90: \newcommand{\Cn}{{\C^n}}
91: \newcommand{\Cns}{(\C^n)^{\text{ss}}}
92: \newcommand{\Cno}{\C^{n+1}}
93: \renewcommand{\cot}{T^*\Cn}
94: \newcommand{\Tk}{T^k}
95: \newcommand{\Tn}{T^n}
96: \newcommand{\Tdrs}{T_{\R}^d\times\Zt}
97: \newcommand{\Tdr}{T_{\R}^d}
98: \newcommand{\Td}{T^d}
99: \newcommand{\Tkd}{(\tk_{\Z})^*}
100: \newcommand{\Tnd}{(\tn_{\Z})^*}
101: \newcommand{\Tdd}{(\td_{\Z})^*}
102: \newcommand{\tk}{\mathfrak{t}^k}
103: \newcommand{\tn}{\mathfrak{t}^n}
104: \newcommand{\td}{\mathfrak{t}^d}
105: \newcommand{\tkd}{(\tk)^*}
106: \newcommand{\tnd}{(\tn)^*}
107: \newcommand{\tdd}{(\td)^*}
108: \newcommand{\Tdso}{\Td\times S^1}
109: \newcommand{\Tnso}{\Tn\times S^1}
110: 
111: \newcommand{\subs}{\subseteq}
112: \newcommand{\sups}{\supseteq}
113: \newcommand{\hookto}{{\hookrightarrow}}
114: \newcommand{\onto}{{\twoheadrightarrow}}
115: \renewcommand{\iff}{\Leftrightarrow}
116: \newcommand{\impl}{\Rightarrow}
117: 
118: \newcommand{\hm}{H^*(M)}
119: \newcommand{\htm}{H^*_{\Td}(M)}
120: \newcommand{\htsm}{H^*_{\Td\times S^1}(M)}
121: \newcommand{\hsm}{H^*_{S^1}(M)}
122: \newcommand{\hr}{H^*_{\Tdr}(M_{\R};\Zt)}
123: \newcommand{\hrs}{H^*_{\Tdr\times\Zt}(M_{\R};\Zt)}
124: \newcommand{\hscomp}{H^*_{\Zt}(\comp;\Zt)}
125: \newcommand{\hcomp}{H^*(\comp;\Zt)}
126: 
127: \newcommand{\becircled}{\mathaccent "7017}
128: \newcommand{\hs}{\hspace{3pt}}
129: \newcommand{\eps}{\varepsilon}
130: \renewcommand{\S}{\sum}
131: \renewcommand{\ss}{\substack}
132: \renewcommand{\bd}{\partial}
133: \newcommand{\gd}{\g^*}
134: \newcommand{\g}{\mathfrak{g}}
135: \newcommand{\half}{\frac{1}{2}}
136: \renewcommand{\mod}{{/\!\!/}}
137: \newcommand{\mmod}{{/\!\!/\!\!/\!\!/}}
138: \newcommand{\mc}{{\mathcal{C}}}
139: \newcommand{\arr}{{\mathcal{H}}}
140: \newcommand{\comp}{\mathcal{M}(\arr)}
141: \newcommand{\bomp}{\Cn\setminus\cup_{i=1}^n H_i^{\C}}
142: \newcommand{\os}{\mathcal{OS}}
143: \newcommand{\ost}{\mathcal{OS}\otimes\Zt}
144: \newcommand{\mhk}{\mu_{\text{HK}}}
145: \renewcommand{\mc}{\mu_{\C}}
146: \newcommand{\mr}{\mu_{\R}}
147: 
148: 
149: \newcommand{\qed}{\hfill \mbox{$\Box$}\medskip\newline}
150: % \newcommand{\qed}{\hfill $\Box$\newline}
151: \newenvironment{proofc}{\noindent {\bf Proof of \ref{coorientation}:}}{\qed\par}
152: \newenvironment{proofp}{\text{ }\newline\noindent 
153: {\bf Proof of \ref{placement}:}}{\qed \par}
154: \newenvironment{proofhtsm}{\text{ }\newline\noindent 
155: {\bf Proof of \ref{htsm}:}}{\qed \par}
156: \newenvironment{proofhsm}{\text{ }\newline\noindent 
157: {\bf Proof of \ref{hsm}:}}{\qed \par}
158: \newenvironment{proof}{\noindent {\bf Proof:}}{\qed \par}
159: \newenvironment{noproof}{\noindent {\bf Proof:}}{\newline}
160: \newenvironment{sketch}{\noindent {\bf Sketch of Proof:}}{\qed \par}
161: 
162: \begin{document}
163: \begin{spacing}{1.1}
164: 
165: \noindent
166: {\LARGE \bf Properties of 
167: the residual circle action \smallskip \\
168: on a hypertoric variety}\bigskip\\
169: {\bf Megumi Harada} \\
170: Department of Mathematics, University of California,
171: Berkeley, CA 94720\smallskip \\
172: {\bf Nicholas Proudfoot } \\
173: Department of Mathematics, University of California,
174: Berkeley, CA 94720
175: \bigskip
176: {\small
177: \begin{quote}
178: \noindent {\em Abstract.}
179: We 
180: consider an orbifold $X$ obtained by a 
181: K\"ahler reduction of $\Cn$,
182: and we define its ``hyperk\"ahler analogue'' $M$
183: as a hyperk\"ahler reduction of $\cot\cong\Hn$ by the same group.
184: In the case where the group is abelian and $X$ is a toric variety,
185: $M$ is a toric hyperk\"ahler orbifold, as defined in
186: \cite{BD}, and further studied in \cite{K1,K2} and \cite{HS}.  
187: The variety $M$ carries a natural action of $S^1$, induced
188: by the scalar action of $S^1$ on the fibers of $\cot$.
189: In this paper we study this action, computing its fixed points
190: and its equivariant cohomology.
191: As an application, we use the associated $\Zt$ action
192: on the real locus of $M$ to compute a deformation of the
193: Orlik-Solomon algebra of a smooth, real hyperplane arrangement $\arr$,
194: depending nontrivially on the affine structure of the arrangement.
195: This deformation is given
196: by the $\Zt$-equivariant cohomology of the complement of the complexification
197: of $\arr$, where $\Zt$ acts by complex conjugation.
198: \end{quote}
199: }
200: \bigskip
201: 
202: In order to construct a toric variety as a K\"ahler quotient of $\Cn$ by a 
203: torus,
204: one begins with the combinatorial data of an arrangement $\arr$ of $n$ cooriented,
205: rational, affine hyperplanes in $\R^d$.
206: The normal vectors to these hyperplanes determine
207: a subtorus $\Tk\subs\Tn$ ($k=n-d$), and the affine structure determines a 
208: value $\a\in\tkd$ at which to reduce, so that we may define $X=\Cn\mod_{\!\!\a}\Tk$.
209: Using the same combinatorial data, one can also construct a {\em hypertoric 
210: variety},\footnote{In 
211: \cite{BD,K1,K2,HS} $M$ is called a ``toric hyperk\"ahler'' variety,
212: but as it is a complex variety that is not toric in
213: the standard sense, we prefer the term ``hypertoric.''}
214: which is defined as the hyperk\"ahler quotient $M = \Hn\mmod_{\!\!(\a,0)}\Tk$ of 
215: $\Hn\cong\cot$ by the
216: induced action of the same subtorus $\Tk\subs\Tn$ \cite{BD}.  It is well known 
217: that the toric variety $X$ does not retain all of the information of $\arr$; indeed, it 
218: depends only on the polyhedron
219: $\Delta$ obtained by intersecting the half-spaces associated to each of the 
220: cooriented hyperplanes. Thus it is always possible to add an extra hyperplane to $\arr$ 
221: without changing $X$.  In contrast, the hypertoric variety $M$ remembers the 
222: number of hyperplanes in $\arr$, but its equivariant diffeomorphism type 
223: depends neither on the coorientations nor on the affine structure of $\arr$ (see Theorem \ref{htm}, 
224: and Lemmas \ref{placement} and \ref{coorientation}).
225: 
226: The purpose of this paper is to study the hamiltonian $S^1$ action on $M$ descending
227: from the scalar action of $S^1$ on the fibers of $\cot$.  This action is sensitive to
228: both the coorientations and the affine structure of $\arr$,
229: even on the level of equivariant cohomology (Section \ref{Konno}).
230: One can recover the toric variety $X$ as the minimum of the $S^1$ moment map,
231: hence the geometric structure of $M$ along with its circle action carries strictly more
232: information than either $X$ or $M$ alone.
233: In Section \ref{coreflow} we give an explicit description of this action
234: when restricted to the core $C$, a deformation retract of
235: $M$ which is a union of projective subvarieties.
236: In Section \ref{Konno} we compute the $S^1$ and $\Tdso$-equivariant cohomologies
237: of $M$, using the full combinatorial data of $\arr$. 
238: 
239: In Section \ref{defos}, we 
240: examine the {\em real locus} $M_{\R}\subs M$, i.e. the fixed point set 
241: of an involution
242: of $M$ that is anti-holomorphic with respect to the first complex structure.
243: By studying the topology of $M_{\R}$, we interpret the results of Section \ref{Konno}
244: in terms of the Orlik-Solomon algebra 
245: $\mathcal{OS}=H^*(\comp)$,
246: where $\comp$
247: is the complement of the complexification of $\arr$.
248: % In particular, we show that $\mathcal{OS}\otimes\Zt$ is isomorphic to a quotient
249: % of $H^*_{\Td}(M;\Zt)$, and that the equivariant cohomology ring $H^*_{\Zt}(\comp;\Zt)$
250: % is isomorphic to the analogous quotient of $H^*_{\Tdso}(M;\Zt)$,
251: % where $\Zt$ acts on $\comp$ by complex conjugation.
252: We show how to interpret Theorem \ref{htsm} as a computation
253: of $H^*_{\Zt}(\comp;\Zt)$,
254: a deformation of the Orlik-Solomon algebra of a smooth, real arrangement
255: that depends nontrivially on the affine structure.\footnote{A more
256: general computation of $H^*_{\Zt}(\comp;\Zt)$, in which $\arr$
257: is not assumed to be simple, rational, or smooth, will appear in \cite{Pr}.}
258: \newline
259: 
260: \noindent {\em Acknowledgments.}
261: We are very grateful to Tam\'as Hausel for introducing us to this problem
262: and answering many questions.
263: We would also like to thank Allen Knutson for guiding our research.
264: 
265: \begin{section}{Hyperk\"ahler reductions}\label{reduction}
266: A hyperk\"ahler manifold is a smooth manifold, 
267: necessarily of real dimension $4n$,
268: which admits
269: three complex structures $J_1,J_2,J_3$
270: %$i,j,k$
271: satisfying the usual quaternionic relations, in a manner compatible
272: with a metric. 
273: Just as in the K\"ahler case, we can define three different
274: symplectic forms on $N$ as follows:
275: % by using the compatibility of the
276: % metric and the complex structures, namely
277: \[
278: \omega_1(v,w) = g(J_{1}v, w), \hs\omega_2(v,w) = g(J_{2}v,w), \hs\omega_3(v,w) =
279: g(J_{3}v,w).
280: \]
281: % As happens in K\"ahler geometry, the integrability of the complex
282: % structures $J_i$ implies the closedness of the two-forms $\omega_i$.
283: Note that the complex-valued two-form \(\omega_2 + i
284: \omega_3\) is non-degenerate and covariant constant, hence closed and
285: {\em holomorphic} with respect to the complex structure $J_1$. 
286: Any hyperk\"aher manifold can therefore be considered as a holomorphic
287: symplectic manifold with complex structure $J_1$, real symplectic
288: form $\omega_{\R} := \omega_1$, and holomorphic symplectic form
289: $\omega_{\C} := \omega_2 + i\omega_3$.
290: This is the point of view that we will adopt in this paper.
291: 
292: % Just as K\"ahler 
293: % reduction is a powerful tool for producing 
294: % examples of K\"ahler manifolds, hyperk\"ahler reduction produces
295: % a plethora of 
296: % many
297: % interesting hyperk\"ahler manifolds. The method
298: % was first introduced 
299: % in a classic paper of 
300: % by Hitchin, Karlhede,
301: % Lindstrom, and Ro\v{c}ek \cite{HKLR},
302: % and is a natural ``hyperk\"ahler'' version of K\"ahler reduction.
303: We will refer to an action of $G$ on a hyperk\"ahler manifold
304: $N$ as {\em hyperhamiltonian}
305: if it is hamiltonian with respect to $\omega_{\R}$ and holomorphic
306: hamiltonian with respect to $\omega_{\C}$, with $G$-equivariant
307: moment map
308: $$\mhk:=\mr\oplus\mc:N\to\gd\oplus\g_{\C}^*.$$ 
309: % where the action of $G$ on $\gd\oplus\g_{\C}^*$ is given by 
310: % the diagonal embedding \(G \to G\times G.\) 
311: 
312: \begin{theorem}{\em\cite{HKLR}}
313: Let \((N^{4n},g)\) be a hyperk\"ahler manifold with real symplectic form
314: $\omega_{\R}$ and holomorphic symplectic form $\omega_{\C}$.
315: Suppose that $N$ is equipped with a hyperhamiltonian action of a 
316: compact Lie group $G$, with moment map
317: $\mhk=\mr\oplus\mc$. 
318: Suppose $\xi = \xi_{\R}\oplus \xi_{\C}$ is a
319: central regular value of $\mhk$. Then there is a unique 
320: hyperk\"ahler structure
321: on the hyperk\"ahler quotient 
322: \(M = N \mmod_{\!\!\xi}G := \mhk^{-1}(\xi)/G\), 
323: with associated symplectic and holomorphic symplectic forms
324: $\omega^{\xi}_{\R}$ and $\omega^{\xi}_{\C}$, such that
325: $\omega^{\xi}_{\R}$ and $\omega^{\xi}_{\C}$ pull back to the restrictions
326: of $\omega_{\R}$ and $\omega_{\C}$ to $\mhk^{-1}(\xi)$.
327: \end{theorem}
328: 
329: % If we fix a complex structure $J_1$, then as mentioned above, the
330: % latter two symplectic forms $\omega_2, \omega_3$ fit together to make
331: % a holomorphic symplectic form $\omega_{\C}$. The latter two moment
332: % maps $\mu_2, \mu_3$ also fit together to give a holomorphic moment map
333: % $\mu_{\C}$. When using this perspective, We can also distinguish
334: % between the moment maps by calling $\mu_1$ the ``real'' moment map,
335: % $\mu_{\R}$. Similarly, we can split the target $\g^{*} \oplus \g^{*}
336: % \oplus \g^{*}$ into components \(\g^{*}_{\R} \oplus \g_{\C}^{*},\) and 
337: % we reduce at values \(\xi = (\xi_{\R}, \xi_{\C}).\) 
338: 
339: % Similarly, it
340: % follows from the Kirwan-Ness theorem that a hyperK\"ahler reduction is
341: % equivalent to a holomorphic symplectic reduction of $M$ by $G_{\C}$,
342: % where the complex group also acts since $G$ preserves the complex
343: % structure $J_1$. Namely, a hyperk\"ahler reduction is by definition
344: % \[
345: % M_{red} := (\mu_{\R}^{-1}(\xi_{\R}) \cap \mu_{\C}^{-1}(\xi_{\C}))/G,
346: % \]
347: % where $G$ is the compact group. 
348: 
349: %BEGIN NEW
350: % Since the hyperk\"ahler moment map $\mhk$ is $G$-equivariant, 
351: If $\xi\in \gd\oplus\g_{\C}^*$ is fixed by the coadjoint action of $G$, the inverse
352: image $\mc^{-1}(\xi_{\C})$ is preserved by $G$,
353: and is a (singular) K\"ahler subvariety with respect to $\omega_{\R}$.
354: Then by 
355: \cite{HL} 
356: (see also \cite[3.2]{Na} and \cite[2.5]{Sj}),
357: we have 
358: $$N\mmod_{\!\!\xi}G = 
359: \mc^{-1}(\xi_{\C})\mod_{\!\!\xi_{\R}}G
360: =\mu_{\C}^{-1}(\xi_{\C})^{\text{ss}}/G_{\C},$$
361: where 
362: $$\mc^{-1}(\xi_{\C})^{\text{ss}} = \{x\in\mc^{-1}(\xi_{\C})\mid Gx\cap\mu_{\R}^{-1}(\xi_{\R})
363: \neq\emptyset\}.$$
364: % is the dense open stratum
365: % of $\mc^{-1}(\xi_{\C})$ in the stratification defined by the gradient
366: % flow of the real-valued Morse-Bott-Kirwan function $|\mr-\xi_{\R}|^2:N\to\R$. 
367: % END NEW
368: 
369: We now specialize to the case where $G$ is a compact Lie group
370: acting linearly on $\Cn$ with moment map $\mu:\Cn \to \gd$, taking $0\in\C^n$
371: to $0\in\gd$.  This action 
372: induces an action of $G$ on the holomorphic cotangent bundle 
373: $\cot\cong\Cn\times(\Cn)^*$.
374: If we choose a bilinear inner product on $\Cn$, 
375: we can coordinatize this representation
376: as $\{(z,w)\mid z,w\in\Cn\}$ with $g (z,w)=(g z,g^{-1} w)$.
377: Choose an identification of $\Hn$ with $\cot$
378: such that the complex structure $J_1$ on $\Hn$ 
379: given by right multiplication by $i$
380: corresponds to the natural complex structure on $\cot$.
381: Then $\cot$ inherits a hyperk\"ahler, and 
382: therefore also a holomorphic symplectic, structure, 
383: with
384: $\omega_{\R}$ given by 
385: adding the standard symplectic structures on $\Cn$ and $(\Cn)^{*}
386: \cong \Cn$, and 
387: $\omega_{\C} = d\eta$, where $\eta$ is the canonical 
388: holomorphic 1-form on $\cot$.  
389: 
390: %Next we note that $\cot$ naturally inherits the structure of a
391: %holomorphic symplectic, and therefore hyperk\"ahler manifold,
392: %with $\omega_{\R}$ given by adding the symplectic forms on $\Cn$ and $(\Cn)^*$,
393: %and $\omega_{\C} = d\eta$, where $\eta$ is the complex valued canonical $1$-form
394: %on $\cot$.
395: % $\omega_{\R} = \frac{i}{2} dz\wedge d\bar{z} - \frac{i}{2} dw\wedge d\bar{w}$
396: %We thus identify $\cot$ with $\Hn$. 
397: 
398: Note that $G$ acts $\H$-linearly on $\cot\cong\Hn$
399: (where $n\times n$ matrices
400: act on the left on $\Hn$, and scalar multiplication by $\H$
401: is on the right), 
402: and does so hyperhamiltonianly
403: with moment map $\mhk=\mc\oplus\mr$,
404: where
405: $$\mr(z,w) = \mu(z)-\mu(w)\hspace{10pt}\text{and}
406: \hspace{10pt}\mc(z,w)(v) = w(\hat{v}_z)$$
407: for $w\in T^*_z\Cn,\hs v\in\g_{\C}$, and $\hat{v}_z$ the element
408: of $T_z\Cn$ induced by $v$.
409: Consider a central regular value $\a\in\gd$ for $\mu$,
410: and suppose that $(\a,0)\in\gd\oplus\gd_{\C}$ is a central regular value for $\mu_{\text{HK}}$.
411: We refer to the hyperk\"ahler reduction $M = \Hn \mmod_{\!\!(\a,0)} G$
412: as the {\em hyperk\"ahler analogue} of the corresponding K\"ahler reduction
413: $X=\Cn\mod_{\!\!\a} G$. 
414: The following proposition is proven for the case 
415: where $G$ is a torus in \cite[7.1]{BD}. 
416: 
417: \begin{proposition}\label{compactification}
418: %Let $M = \Hn \mmod_{\!\!(\a,0)} G$ 
419: %be the hyperk\"ahler reduction of $\Hn$ by $G$
420: %at a central regular value $\left(\a,0\right)\in\gd\oplus\gd_{\C}$.
421: The cotangent bundle $T^*X$ is isomorphic to an open subset of $M$.
422: \end{proposition}
423: 
424: \begin{proof}
425: Let $Y = \{(z,w)\in\mc^{-1}(0)^{\text{ss}}\mid
426: z \in (\Cn)^{\text{ss}}\}$,
427: where we ask $z$ to be semistable with respect to $\a$
428: for the action of
429: $G_{\C}$ on $\Cn$, so that $X\cong (\Cn)^{\text{ss}}/G_{\C}$.
430: Let $[z]$ denote the element of $X$ represented by $z$.
431: The tangent space $T_{[z]}X$ is equal to the quotient of $T_{z}\Cn$
432: by the tangent space to the $G_{\C}$ orbit through $z$,
433: hence
434: $$T^*_{[z]}X\cong\{w\in T^*_{[z]}\Cn\mid w(\hat{v}_z)=0\text{ for all }
435: v\in\g_{\C}\} = \{w\in(\Cn)^*\mid\mc(z,w)=0\}.$$
436: Then $$T^*X \cong \{(z,w)\mid z\in(\Cn)^{\text{ss}}\text{ and }
437: \mc(z,w)=0\}/G_{\C} = Y/G_{\C}$$
438: is an open subset of $M$.
439: \end{proof}
440: 
441: Consider the action
442: of $S^1$ on $\Hn\cong T^*\Cn$ given by 
443: ``rotating the fibers'' of the cotangent
444: bundle,  
445: given explicitly by $\tau(z,w) = (z,\tau w)$. 
446: This action is hamiltonian with respect to the real symplectic
447: structure $\omega_{\R}$ with moment map $\Phi(z,w) = \half |w|^2$.
448: Because it commutes with the action of $G$,
449: the action descends to a hamiltonian action on $M$, where we will
450: still denote the moment map by $\Phi$.
451: Since $S^1$ acts trivially on $z$, and by scalars on $w$,
452: it does {\em not} preserve the complex symplectic form
453: $\omega_{\C}(z,w) = dw\wedge dz$,
454: and does not act $\H$-linearly.
455: % Further, this $S^1$ does {\em not} 
456: % act on $\cot \cong \H^n$ $\H$-linearly, i.e. as elements of 
457: % $M(n,\H)$. 
458: 
459: \begin{proposition}\label{proper}
460: If the original moment map $\mu:\Cn\to\gd$ is proper,
461: then so is $\Phi:M\to\R$.
462: \end{proposition}
463: 
464: \begin{proof}
465: We would like to show that $\Phi^{-1}[0,R]$ is compact
466: for any $R$.  Since 
467: $$\Phi^{-1}[0,R] = \{(z,w)\mid \mr(z,w)=\a,\hs
468: \mc(z,w)=0,\hs\Phi(z,w)\leq R\}\big/G$$ and $G$ is compact,
469: it is sufficient to show that 
470: the set $\{(z,w)\mid \mr(z,w)=\a,\hs\Phi(z,w)\leq R\}$
471: is compact.  Since $\mr(z,w)=\mu(z)-\mu(w)$,
472: this set is a closed subset
473: of $$\mu^{-1}\left\{\a+\mu(w) \,\big|\, \half |w|^2\leq R\right\}
474: \times\left\{w \,\big|\,
475: \half |w|^2 \leq R\right\},$$
476: which is compact by the properness of $\mu$.
477: \end{proof}
478: 
479: % It is important to note that, while the set of non-regular values of $\mu$
480: % has codimension $1$, the set of non-regular values of $\mhk$
481: % has codimension $3$, hence the set of regular values is simply-connected.
482: % Connectivity tells us that the diffeomorphism type of $M$ is independent of $\a$,
483: % and simple-connectivity tells us that the cohomology rings of $M$ and $M'$
484: % are naturally isomorphic for any regular values $\a, \a'\in\gd$.
485: % It is not true, however, that the regular values $(\a,0)$ and $(\a',0)$ of $\mhk$
486: % can be connected by a path of regular values of the form $(\b,0)$.
487: % This means that while the cohomology of $M$ is independent of $\a$,
488: % the $S^1$ equivariant cohomology is not.  In Section \ref{s1}, 
489: % we will compute this equivariant cohomology ring in the case where $G$
490: % is abelian, and $X = \Cn\mod G$ is a toric variety.  
491: 
492: In the case where $G$ is abelian and $X$ is a nonempty toric 
493: variety, properness of $\mu$ (and therefore of $\Phi$) is equivalent 
494: to compactness of $X$. 
495: \end{section}
496: 
497: \begin{section}{Hypertoric varieties}\label{ht}
498: In this section we restrict our attention to 
499: {\em hypertoric}
500: varieties,
501: which are the hyperk\"ahler analogues
502: of toric varieties in the sense of
503: Section \ref{reduction}.  
504: % We
505: % will follow the exposition of \cite{BD,K1,K2}, but some of our notation
506: % will differ.
507: We begin with the full $n$-dimensional torus $\Tn$ acting on $\Cn$,
508: and the induced action on $\Hn\cong\cot$
509: given by $t(z,w) = (tz, t^{-1} w)$.
510: Let \(\{a_i\}_{1 \leq i \leq n}\) be nonzero primitive integer vectors in
511: \(\liet^d \cong \R^d\) defining a map \(\beta: \liet^n \longrightarrow
512: \liet^d\) by \(\e_i \mapsto a_i,\) where
513: $\{\e_i\}$ is the standard basis for \(\liet^n \cong \R^n\), dual to $\{u_i\}$.
514: This map fits into an exact sequence
515: \[
516: 0 \longrightarrow \liet^k \stackrel{\iota}{\longrightarrow} \liet^n
517: \stackrel{\beta}{\longrightarrow} \liet^d\longrightarrow 0,
518: \]
519: where \(\liet^k := {\mathrm ker}(\beta).\) Exponentiating, we
520: get the exact sequence
521: \[
522: 0 \longrightarrow T^k \stackrel{\iota}{\longrightarrow} T^n
523: \stackrel{\beta}{\longrightarrow} T^d \longrightarrow 0,
524: \]
525: whereas by dualizing, we get
526: \[
527: 0 \longrightarrow (\liet^d)^{*} \stackrel{\beta^{*}}{\longrightarrow}
528: (\liet^n)^{*} \stackrel{\iota^{*}}{\longrightarrow} (\liet^k)^{*} \longrightarrow 0,
529: \]
530: where we abuse notation by using $\iota$ and $\beta$
531: to denote maps on the level of groups as well as on the level of algebras.
532: Note that $T^k$ is
533: connected if and only if the vectors $\{a_1,\ldots,a_n\}$ span
534: $\mathfrak{t}^d$ over the integers.
535: 
536: Consider the restriction of the action of $T^n$ on $\HH^n$ to the
537: subgroup $T^k$. This action is hyperhamiltonian with hyperk\"ahler moment map
538: $$\bar{\mu}_{\R}(z,w) = \iota^* \left(\frac{1}{2} \sum_{i=1}^n (|z_i|^2 - |w_i|^2) u_i
539: \right) \hspace{15pt}
540: \text{and}\hspace{15pt}\bar{\mu}_{\C}(z,w) = \iota^* \left(\sum_{i=1}^n (z_i w_i)u_i\right),$$
541: where $\{u_i\}$ is the standard basis in \(\tnd \cong \R^n.\)
542: In contrast with the K\"ahler situation, the hyperk\"ahler moment
543: map is surjective onto $\tnd\oplus(\tn_{\C})^*$.
544: 
545: We denote by 
546: % \(M(\bar{a}, \alpha)\)
547: $M$ the hyperk\"ahler
548: reduction of $\HH^n$ by the subtorus $T^k$ 
549: % (determined by the \(\bar{a} = (a_i)\)) 
550: at \((\alpha, 0)
551: \in (\liet^k)^{*}\oplus (\tk_{\C})^*,\) which is the 
552: hyperk\"ahler analogue of the K\"ahler toric variety $X = \Cn\mod_{\!\!\a}T^k$. 
553: Choose a lift $\tilde\a\in\tnd$ of $\a$ along $\iota^*$.
554: Then $M$ has a natural residual
555: action of $T^d$ with hyperk\"ahler moment map \(\mhk = \mr\oplus\mc.\)
556: Note that the choice of subtorus $\Tk\subs\Tn$ is equivalent
557: to choosing a central arrangement of cooriented hyperplanes in $\tdd$,
558: where the $i^{\text{th}}$ hyperplane is the annihilator
559: of $a_i\in\td$.  (The coorientation comes from the fact that we know
560: for which $x$ we have $\< x,a_i\> > 0$.)
561: The choice of $\tilde\a$ corresponds to an 
562: affinization $\arr$ of this arrangement,
563: where the $i^{\text{th}}$ hyperplane is 
564: $$H_i = \{x\in\tdd\mid\< x,a_i\>=\< -\tilde\a,\e_i\>\}.$$
565: Changing $\tilde\a$ by an element $c\in\tdd$ has the effect of translating $\arr$ by $c$,
566: and adding $c$ to the residual moment map $\mr$.
567: In order to record the information about coorientations,
568: we define the half-spaces 
569: \begin{equation}\label{FG}
570: F_i = \{x\in\tdd\mid\< x,a_i\>\geq\< -\tilde\a,\e_i\>\}
571: \hspace{15pt}\text{and}\hspace{15pt} 
572: G_i = \{x\in\tdd\mid\< x,a_i\>\leq\< -\tilde\a,\e_i\>\},
573: \end{equation}
574: which intersect in the hyperplane $H_i$.
575: Our convention will be to draw pictures, as in Figure 1,
576: in which we specify the coorientations
577: of the hyperplanes by shading the polyhedron $\Delta=\cap_{i=1}^n F_i$
578: (which works as long as $\Delta\neq\emptyset$).
579: Note that the K\"ahler variety $X$ is precisely the 
580: K\"ahler toric variety determined by $\Delta$. 
581: 
582: \begin{figure}[h]
583: \centerline{\epsfig{figure=htv2.eps}}
584: \caption{A hypertoric variety of real dimension 8 obtained
585: by reducing $\H^4$ by $T^2$.}
586: \end{figure}
587: 
588: % Bielawski and Dancer \cite{BD} show that 
589: % $M(\bar{a}, \alpha)$ 
590: The variety $M$ is an orbifold if and only if 
591: $\arr$ is simple, i.e. if and only if
592: every subset of $m$ hyperplanes intersect in codimension $m$ \cite[3.2]{BD}.
593: Furthermore, $M$ is smooth if and only if whenever
594: some subset of $d$ hyperplanes $\{H_i\}$ has non-empty intersection, 
595: the corresponding
596: vectors $\{a_i\}$ form a $\Z$-basis for \(\Z^d \subseteq \liet^d.\)
597: In this case we will refer to the arrangement itself as {\em smooth}.
598: We will always assume that $\arr$ is simple, and at times we will also assume that it is smooth.
599: % We will always assume that $\arr$ 
600: % is smooth and simple, and when we discuss
601: % cohomology in Sections \ref{Konno} and \ref{s1},
602: % we will also assume that
603: 
604: The hyperplanes $\{H_i\}$
605: divide \((\liet^d)^{*} \cong \R^d\) into a finite family of closed,
606: convex polyhedra $$\Delta_A = (\cap_{i\in A}F_i)\cap(\cap_{i\notin A}G_i),$$
607: indexed by subsets $A\subs\{1,\ldots,n\}$.
608: % , i.e. $\Delta_A$
609: % is
610: % obtained from $\Delta$ by flipping the coorientations of the hyperplanes corresponding
611: % to the elements of $A$. 
612: % For example, $\Delta_{\emptyset} = \Delta$ is the shaded region
613: % in Figure 1, and $\Delta_{\{2\}}$ is the triangle above $\Delta$.
614: Consider the subset $I = 
615: \{A\subs\{1,\ldots,n\}\mid \Delta_A\text{  bounded}\}$
616: of the power set of $\{1,\ldots,n\}$.  
617: For each $A\subs\{1,\ldots,n\}$, let $$M_A = \mr^{-1}(\Delta_A)\cap\mc^{-1}(0).$$
618: The K\"ahler submanifold
619: $\left(M_A,\omega_{\R}|_{M_A}\right)$ of $\left(M,\omega_{\R}\right)$
620: is $d$-dimensional and invariant under the action of $\Td$,
621: and is therefore $\Td$-equivariantly isomorphic
622: to the K\"ahler toric variety determined by $\Delta_A$ \cite[6.5]{BD}.
623: We define the {\em core} $C$ and {\em extended core} $D$
624: of a hypertoric variety by setting $$C = \cup_{A\in I}M_A\hs\hs\text{ and }\hs\hs
625: D = \cup_{A}{M_A} = \mc^{-1}(0) =
626: \{[z,w]\mid z_iw_i=0\text{  for all  }i\},$$
627: where $[z,w]$
628: denotes the $T^k$-equivalence class in $M$ of the element
629: $(z,w)\in \bar{\mu}_{\text{HK}}^{-1}(\a,0)$.
630: Bielawski and Dancer \cite{BD} show that
631: $C$ and $D$ are each $\Td$-equivariant
632: deformation retracts of $M$.  See Corollary \ref{unstable} for a Morse theoretic proof.
633: 
634: We take a minute to discuss the differences between
635: the combinatorial data
636: determining a toric variety $X = \Cn\mod_{\!\!\a}\Tk$ and its hypertoric
637: analogue $M = \Hn\mmod_{\!\!(\a,0)}\Tk$.
638: Each is determined by $\arr$,
639: a simple, cooriented, affine arrangement of $n$ hyperplanes in $\tdd$,
640: defined up to simultaneous translation.
641: The toric variety $X$ is in fact determined by less information than
642: this; it depends only on the polyhedron $\Delta = \cap_{i=1}^n F_i$.
643: Thus if the last hyperplane $H_n$ has the property that
644: $\cap_{i=1}^{n-1}F_i\subs F_n$, then this hyperplane is superfluous to $X$.
645: This is not the case for $M$, which means that it is slightly
646: misleading to call $M$ the hyperk\"ahler analogue of $X$;
647: more precisely, it is the hyperk\"ahler analogue of
648: {\em a given presentation} of $X$ as a K\"ahler reduction of $\Cn$.
649: On the other hand, the $\Td$-equivariant diffeomorphism type of $M$ also 
650: does not depend on all of the information of $\arr$, as evidenced by the two following results.
651: 
652: \begin{lemma}\label{placement}
653: The hypertoric varieties \(M_{\a} = \Hn \mmod_{\!\!(\a,0)}\Tk\) 
654: and $M_{\a'}= \Hn \mmod_{\!\!(\a',0)}\Tk$ are $\Td$-equivariantly
655: diffeomorphic, and their
656: cohomology rings can be naturally identified.
657: \end{lemma}
658: 
659: \begin{lemma}\label{coorientation}
660: The hypertoric variety $M$ does not depend on the coorientations
661: of the hyperplanes $\{H_i\}$.
662: \end{lemma}
663: 
664: This means that, unlike that of $X$, 
665: the $\Td$-equivariant diffeomorphism type of $M$ depends only on the unoriented
666: central arrangement underlying $\arr$.
667: A weaker version of Lemma \ref{placement}, involving
668: the (nonequivariant) homeomorphism type of $M$, appears in \cite{BD}.
669: 
670: \begin{proofp}
671: The set of nonregular values for \(\bar{\mu}_{\text{HK}}\) 
672: has codimension 3 inside of \(\tdd
673: \oplus (\td_{\C})^*.\) This tells us that the set of regular values 
674: is simply connected, and we can
675: choose a path connecting any two regular values \((\alpha,0)\)
676: and \((\alpha',0),\) 
677: unique up to homotopy.
678: 
679: Since the moment map \(\bar{\mu}_{\text{HK}}\)
680:  is not proper, we must take some care in showing that
681: two fibers are diffeomorphic. To this end, we note that the
682: norm-square function \(\psi(z,w) = \|z\|^2 + \|w\|^2\) is $T^n$-invariant
683: and proper on $\Hn$. 
684: Let $\H_{reg}^n$ denote the open submanifold of $\Hn$ consisting of the
685: preimages of the regular values of \(\bar{\mu}_{\text{HK}}.\) 
686: By a direct computation, it is easy
687: to see that the kernels of $d\psi$ and \(d(\bar{\mu}_{\text{HK}})\) intersect
688: transversely at any point \(p \in \H_{reg}^n.\) Using the standard
689: $T^n$-invariant metric on $\Hn$, we define an Ehresmann connection
690: on $\H_{reg}^n$ with respect to \(\bar{\mu}_{\text{HK}}\) 
691: such that the horizontal subspaces
692: are contained in the kernel of $d\psi$.
693: %Note that this connection can be extended to an Ehresmann
694: %connection with respect to \(\iota^{*} \bar{\mu}_{\R} \oplus
695: %\iota_{\C}^{*} \bar{\mu}_{\C} \oplus \psi.\) 
696: 
697: This connection allows us to lift a path connecting the two
698: regular values to a horizontal vector field on its preimage in
699: $\H_{reg}^n$. Since the horizontal subspaces are tangent to
700: the kernel of $d\psi$, the flow preserves level sets of $\psi$. Note that the
701: function \[\bar{\mu}_{\text{HK}} \oplus \psi: \Hn \to \tdd \oplus (\td_{\C})^{*} \oplus
702: \R\] {\it is} proper. By a theorem of Ehresmann \cite[8.12]{BJ}, the
703: properness of this map
704: implies that the flow of this vector field exists for all time, and
705: identifies the inverse image of $(\alpha,0)$ with that of
706: $(\alpha',0)$. Since the metric, $\psi$, and \(\bar{\mu}_{\text{HK}}\) are
707: all $T^n$-invariant, the Ehresmann connection is also $T^n$-invariant,
708: therefore the diffeomorphism identifying the fibers
709: is $T^n$-equivariant, making 
710: the reduced spaces are $T^d$-equivariantly diffeomorphic. 
711: \end{proofp}
712: 
713: \vspace{-\baselineskip}
714: \begin{proofc}
715: It suffices to consider the case when we change the orientation of a
716: single hyperplane within the arrangement.  Changing the coorientation
717: of a hyperplane $H_l$ is equivalent to defining a new map
718: $\beta':\tn\to\td$, with $\beta'(\e_i)=a_i$ for $i\neq l$, and $-a_i$ for $i=l$.
719: This map exponentiates to a map $\beta':\Tn\to\Td$,
720: and we want to show that the hyperk\"ahler variety obtained
721: by reducing $\Hn$ by the torus $\ker(\beta')$ is isomorphic to $M$,
722: which is obtained by reducing $\Hn$ by the torus $\Tk=\ker(\beta)$.
723: To see this, note that $\ker(\beta')$ and $\ker(\beta)$
724: are conjugate inside of $\operatorname M(n,\H)$ by the element
725: $(1,\ldots,1,j,1,\ldots,1)\in \operatorname 
726: M(1,\H)^n\subs \operatorname M(n,\H)$,
727: where the $j$ appears in the $l^{\text{th}}$ slot. 
728: \end{proofc}
729: 
730: \vspace{-\baselineskip}
731: \begin{example}\label{same}
732: The three cooriented arrangements of Figure 2 all specify
733: the same hyperk\"ahler variety $M$ up to equivariant
734: diffeomorphism.  The first has $X\cong \mathbb{F}_1$ (the
735: first Hirzebruch surface) and the second and the third have $X\cong\C P^2$.
736: Note that if we flipped the coorientation of $H_3$ in Figure 2(a)
737: or 2(c), then
738: we would get a non-compact
739: $X\cong\tilde{\C}^2$, the blow-up of $\C^2$
740: at a point.  If we flipped the coorientation of $H_3$ in
741: Figure 2(b), then $X$ would be empty.
742: We make no assumptions about $X$ in this section.
743: 
744: \begin{figure}[h]
745: \centerline{\epsfig{figure=htv1.eps}}
746: \caption{Three arrangements
747: related by flipping coorientations and translating hyperplanes.}
748: \end{figure}
749: \end{example}
750: 
751: The purpose of this paper is to study not just the topology
752: of $M$, but the topology of $M$ along with the natural hamiltonian
753: $S^1$ action defined in Section \ref{reduction}.
754: In order to define this $S^1$ action, it is necessary that
755: we reduce at a regular value of the form $(\a,0)\in\tdd\oplus(\liet_{\C}^d)^*$,
756: and although the set of regular values of $\bar{\mu}_{\text{HK}}$ is simply connected,
757: the set of regular values of the form $(\a,0)$ is not even connected.
758: Furthermore, 
759: left multiplication by the diagonal matrix $(1,\ldots,1,j,1,\ldots,1) \in 
760: \operatorname U(n,\H)$ is not
761: an $S^1$-equivariant automorphism of $\Hn$, therefore the geometric
762: structure of $M$ along with a circle action may depend nontrivially
763: both on the affine structure
764: and the coorientations of the arrangement $\arr$.
765: Indeed it must, because we can recover $X$ from $M$ by taking the minimum
766: $\Phi^{-1}(0)$ of the $S^1$ moment map $\Phi:M\to\R$.
767: In this sense, the structure of a hypertoric variety $M$
768: along with a circle action is the universal geometric object from which
769: both $M$ and $X$ can be recovered.
770: % In Section \ref{s1}, we will see that
771: % the $S^1$-equivariant cohomology ring of $M$ is sensitive
772: % to the coorientations as well as the affine structure of $\arr$.
773: \end{section}
774: 
775: \begin{section}{Gradient flow on the core}\label{coreflow}
776: Although $S^1$ does not act on $M$ as a subtorus of $\Td$, we show
777: below that when restricted to any single 
778: component $M_A$ of the extended core, $S^1$
779: {\em does} act as a subtorus of $\Td$, with the subtorus depending
780: combinatorially on $A$.  This will allow us to give a combinatorial
781: analysis of the gradient flow of $\Phi$ on the extended core.
782: 
783: \begin{lemma}\label{sides}
784: Let $x$ be an element of $\tdd$, and consider a point $[z,w]\in\mr^{-1}(x)\cap D$.
785: We have $x \in F_i$ if and only if $w_i=0$, and $x\in G_i$
786: if and only if $z_i=0$.
787: \end{lemma}
788: 
789: \begin{proof}
790: The fact that $[z,w]\in D$ tells us that $z_iw_i=0$.
791: Then 
792: \begin{eqnarray*}
793: x\in F_i & \iff & \< -\tilde\a,\e_i\>\leq\<\mr[z,w],
794: 	           a_i\>  =\<\bar\mr(z,w)-\tilde\a,\e_i\>  \\ 
795:         & \iff  &  \half|z_i|^2-\half|w_i|^2=\<\bar\mr(z,w),
796: 		    \e_i\>\geq 0.                                  \\
797: \end{eqnarray*}
798: Since $z_iw_i=0$, this is equivalent to the condition $w_i=0$.
799: The second half of the lemma follows similarly.
800: \end{proof}
801: 
802: On the suborbifold $M_A\subs D\subs M$ we have $z_i = 0$ for all $i\in A$
803: and $w_i = 0$ for all $i\notin A$, therefore for $\tau\in S^1$
804: and $[z,w]\in M_A$,
805: $$\tau[z,w]=[z,\tau w] = 
806: [\tau_1 z_1,\ldots \tau_n z_n, \tau_1^{-1}w_1,\ldots \tau_n^{-1}w_n], \text{ where }
807: \tau_i =
808: \begin{cases}
809: \tau^{-1} & \text{if }i\in A,\\
810: 1 & \text{if }i\notin A.
811: \end{cases}
812: $$
813: In other words, the $S^1$ action on $M_A$
814: is given by the one dimensional subtorus 
815: $(\tau_1, \ldots, \tau_n)$
816: of the original torus $T^n$,
817: hence the moment map $\Phi|_{M_A}$ is given (up to
818: an additive constant) by 
819: $$\Phi[z,w] = \left<\mr[z,w],\sum_{i\in A}a_i\right>.$$
820: This formula allows us to compute the fixed point sets of
821: the $S^1$ action.
822: Since $S^1$ acts freely on \((\liet_{\C}^d)^{*} \backslash \{0\}\)
823: and \(\mc: M \to (\liet_{\C}^d)^{*}\) is $S^1$-equivariant,
824: we must have $M^{S^1} \subseteq \mc^{-1}(0) = D$.
825: For any subset $B\subs\{1,\ldots,n\}$, let $M_A^B$ be the toric subvariety
826: of $M_A$ defined by the conditions $z_i=w_i=0$ for all $i\in B$.
827: Geometrically, $M_A^B$ is defined by the (possibly empty) face
828: $\cap_{i\in B}H_i \cap \Delta_A$ of the polyhedron $\Delta_A$.
829: 
830: \begin{proposition}\label{fixed}
831: The fixed point set of the action of $S^1$ on $M_A$
832: is the union of those toric subvarieties $M_A^B$ such that
833: $\sum_{i\in A}a_i\in \liet^d_B:=\operatorname{Span}_{j\in B}a_j$.
834: \end{proposition}
835: 
836: \begin{proof}
837: The moment map $\Phi|_{M_A^B}$ will be constant if and only if
838: $\sum_{i\in A}a_i$ is perpendicular to 
839: $\ker\left(\tdd\onto(\liet^d_B)^*\right)$,
840: i.e. if $\sum_{i\in A}a_i$ lies in the kernel of the projection
841: $\td\onto\td/\liet^d_B$.
842: \end{proof}
843: 
844: \vspace{-\baselineskip}
845: \begin{corollary}
846: Every vertex $v\in\tdd$ of the polyhedral complex defined by $\arr$
847: is the image of an $S^1$-fixed point in $M$.
848: Every component of $M^{S^1}$ has dimension less than or equal to $d$,
849: and the only component of dimension $d$ is $M_{\emptyset}=X=\Phi^{-1}(0)$. 
850: \end{corollary}
851: 
852: For any point $p\in M^{S^1}$, the {\em stable orbifold} 
853: $S(p)$ at $p$ is defined to be the set of $x\in M$ such
854: that $x$ approaches $p$ when flowing along the vector field 
855: $-\operatorname{grad}(\Phi)$,
856: and the {\em unstable orbifold} $U(p)$ at $p$ is defined to be the stable orbifold with respect
857: to the function $-\Phi$.
858: For any suborbifold $Y\subs M^{S^1}$, the unstable orbifold $U(Y)$ at $Y$ is defined
859: to be the union of $U(y)$ for all $y\in Y$.  In general, for $y\in Y$, we have the identity
860: $\dim_{\R} U(Y) + \dim_{\R} S(y) = 4d$.
861: 
862: Let $Y\subs M^{S^1}$ be a component of the fixed point set of $M$.
863: Let $v\in\tdd$ be a vertex in the polehedron $\mr(Y)$, and let $y$
864: be the unique preimage of $v$ in $Y$.
865: 
866: \begin{proposition}
867: The unstable orbifold $U(Y)$ is a complex suborbifold of complex 
868: dimension at most $d$, contained in the core $C\subs M$.
869: If $\arr$ is smooth at $y$, then $\dim_{\C}U(Y)=d$, and the closure of $U(Y)$
870: is an irreducible component of $C$.
871: \end{proposition}
872: 
873: \begin{proof}
874: For simplicity, we will assume that $v=\cap_{j=1}^dH_j$.
875: For all $l\in\otd$, let $b_l\in\td_{\Z}$ be the smallest integer vector
876: such that $\left<a_j,b_l\>=0$ for $j\neq l$ and $\left<a_l,b_l\> >0$.
877: Geometrically, $b_l$ is the primitive integer vector on the line
878: $\cap_{j\neq l}H_j$ pointing in the direction of $\Delta$.
879: Note that $M$ is smooth at the $T^d$-fixed point above $v$
880: if and only if $\left<a_l,b_l\>=1$ for all $l\in\otd$.
881: Let $R_l\subs\tdd$ be the ray eminating from $v$ in the direction of $b_l$,
882: and ending before it hits another vertex.  Let $Q_l$ be the analogous ray in the opposite
883: direction.
884: 
885: Let $\Delta_A$ be a region (not necessarily bounded) of the polyhedral 
886: complex defined by $\arr$ adjacent to $R_l$.
887: The preimage $\mr^{-1}(R_l)\cap D$ of $R_l$ in $D$ is a complex line,
888: and it is contained in the unstable
889: orbifold at $U(Y)$ if and only if $\left<b_l,\sum_{i\in A}a_i\right>\geq 0$.
890: If $\left<b_l,\sum_{i\in A}a_i\right> <0$, it is contained in the stable orbifold $S(y)$.
891: The preimage $\mr^{-1}(Q_l)\cap D$ of $Q_l$ in $D$ is also a complex line,
892: contained in the unstable
893: orbifold $U(Y)$ if and only if $\<-b_l,a_l+\sum_{i\in A}a_i\>\geq 0$,
894: and otherwise in $S(y)$.
895: Since $\<b_l,\sum_{i\in A}a_i\>+\<-b_l,a_l+\sum_{i\in A}a_i\>
896: =-\left<a_l,b_l\> <0$, at most one of these two directions can be unstable.
897: In the smooth case, $\left<a_l,b_l\>=1$ for all $l$, and exactly one of
898: the two directions is unstable.
899: 
900: Consider the polytope $\Delta_v$ incident to $v$ and characterized by the property that its
901: edges at the vertex $v$ are exactly the unstable directions. 
902: The toric variety $X_{\Delta_v}\subs D$ is contained
903: in the closure of $U(Y)$, and a dimension count tells us that this containment is an equality.
904: In the smooth case, $\Delta_v$ is $d$-dimensional, and $X_{\Delta_v}$ is a component of the core.
905: \end{proof}
906: 
907: Note that, even in the smooth case, it is not necessarily the case that the
908: $R_l$ direction is stable and the $Q_l$ direction is unstable.  See, for example,
909: the vertex $v = H_1\cap H_2$ in Figure 2(c).
910: 
911: \begin{corollary}
912: There is a natural injection from the set of bounded regions 
913: $\{\Delta_A\mid A\in I\}$
914: to the set of connected components of $M^{S^1}$.
915: If $\arr$ is smooth, this map is a bijection.
916: \end{corollary}
917: 
918: \begin{proof}
919: To each $A\in I$, we associate the fixed subvariety $M_A^B$ corresponding
920: to the face of $\Delta_A$ on which the linear functional $\sum_{i\in A}a_i$
921: is minimized, so that $M_A$ is the closure of
922: $U(M_A^B)$.  If $\arr$ is smooth, then every connected component of the fixed point
923: set will have a component of the core as its closed unstable orbifold.
924: \end{proof}
925: 
926: \vspace{-\baselineskip}
927: \begin{corollary}\label{unstable}
928: The core of $M$ is equal to the union of the unstable orbifolds of
929: the connected components of $M^{S^1}$, hence $C$ is a $\Tdso$-equivariant
930: deformation retract of $M$.
931: \end{corollary}
932: 
933: \begin{example}
934: In Figure 3, representing a reduction of $\H^5$ by $T^3$,
935: we choose a metric on $(\liet^2)^*$ in order
936: to draw the linear functional $\sum_{i\in A}a_i$ as a vector
937: in each region $\Delta_A$.  We see that $M^{S^1}$ has three components,
938: one of them $X\cong\mathbb F_1$, one of them a projective line, with another $\mathbb F_1$
939: as its unstable manifold, and one of them a point, with a $\C P^2$ as its unstable manifold.
940: 
941: \begin{figure}[h]
942: \centerline{\epsfig{figure=flow.eps}}
943: \caption{The gradient flow of $\Phi:M\to\R$.}
944: \end{figure}
945: \end{example}
946: 
947: \begin{example}
948: The hypertoric variety represented by Figure 4 has a fixed point set with four
949: connected components (three points and a $\C P^2$), but only three components in its core.
950: This phenomenon can be blamed on the orbifold point represented by the
951: intersection of $H_3$ and $H_4$, which has a one-dimensional unstable orbifold.
952: 
953: \begin{figure}[h]
954: \centerline{\epsfig{figure=orbi.eps}}
955: \caption{A singular example.}
956: \end{figure}
957: \end{example}
958: \end{section}
959: 
960: \begin{section}{Equivariant cohomology}\label{Konno}
961: In this section we extend Konno's computations
962: of the ordinary and $\Td$-equivariant cohomologies of $M$
963: to the $S^1$-equivariant setting.
964: We follow Konno's approach of restricting to the smooth case
965: to simplify arguments involving line bundles on $M$.
966: Hausel and Sturmfels, however, prove
967: theorems analogous to \ref{htm} and \ref{hm} with rational coefficients 
968: in the orbifold case, and Theorems \ref{htsm} and \ref{hsm}
969: extend to this setting as well (see Remark \ref{lawrence}).
970: 
971: \begin{theorem}\label{htm}{\em\cite{K2}}
972: The $\Td$-equivariant cohomology ring of a smooth 
973: hypertoric variety $M$ is given by
974: $$\htm = \Z[u_1,\ldots,u_n]\bigg/
975: \left<\prod_{i\in S}u_i\hs\bigg{\vert}\hs\bigcap_{i\in S} H_i = \emptyset \right>.$$
976: \end{theorem}
977: 
978: \begin{remark}
979: This is precisely
980: the Stanley-Reisner ring of
981: the unoriented matroid determined by the arrangement $\arr$ \cite{HS}.
982: \end{remark}
983: 
984: Just as the cohomology of a toric variety is obtained
985: from the equivariant cohomology by introducing linear relations
986: that generate $\ker\iota^*=\left(\ker\beta\right)^{\bot}$, the same
987: is true for hypertoric varieties:
988: 
989: \begin{theorem}\label{hm}{\em\cite{K1}}
990: The ordinary cohomology ring of a smooth hypertoric variety $M$ is given by
991: $$\hm = \htm \big/ \<\Sigma a_iu_i\in \ker\iota^*\>.$$
992: \end{theorem}
993: 
994: The rest of this section will be devoted to the proof of the following two theorems.
995: % Given any set $S\subs\{1,\ldots,n\}$ such that $\cap_{i\in S} H_i = \emptyset$,
996: % there exists a splitting (unique if $S$ is minimal) $S = S_1\sqcup S_2$
997: % such that
998: % $$\big(\cap_{i\in S_1}G_i\big)\cap
999: % \big(\cap_{j\in S_2}F_j\big) = \emptyset.$$
1000: 
1001: \begin{theorem}\label{htsm}
1002: Let $M$ be the hypertoric variety corresponding to a
1003: smooth, cooriented arrangement $\arr$.
1004: Given any minimal set $S\subs \{1,\ldots,n\}$ such that 
1005: $\cap_{i\in S} H_i = \emptyset$, let $S = S_1\sqcup S_2$
1006: be the unique splitting of $S$ such that
1007: $\big(\cap_{i\in S_1}G_i\big)\cap
1008: \big(\cap_{j\in S_2}F_j\big) = \emptyset$ (see \eqref{FG}).
1009: Then the $\Tdso$-equivariant cohomology of $M$ is given by
1010: $$\htsm\cong\Z[u_1,\ldots,u_n,x]\bigg/
1011: \< \prod_{i\in S_1}u_i\times\prod_{j\in S_2}(x-u_j)
1012: \hs\bigg{\vert}\hs\bigcap_{i\in S} H_i = \emptyset\>.$$
1013: \end{theorem}
1014: 
1015: \begin{theorem}\label{hsm}
1016: In the notation of Theorem \ref{htsm},
1017: the $S^1$-equivariant cohomology ring
1018: of $M$ is given by
1019: $$\hsm \cong \htsm \big/ \<\Sigma a_iu_i\in \ker\iota^*\>.$$
1020: \end{theorem}
1021: 
1022: \begin{remark}
1023: Konno observes that the quotient map from the abstract
1024: polynomial ring $\Z[u_1,\ldots,u_n]\to\htm$ is precisely the $\Td$-equivariant
1025: Kirwan map $$\ktd:H^*_{\Tn}(\cot)\to\htm$$
1026: induced by the inclusion
1027: $\mu^{-1}(\a,0)\hookto\cot$.  
1028: Likewise, the map from $\Z[u_1,\ldots,u_n]/\ker\iota^*$ to $H^*(M)$
1029: is the ordinary Kirwan map
1030: $$\k:H^*_{\Tk}(\cot)\to H^*(M).$$
1031: The analogous maps for K\"ahler reductions
1032: are known to always be surjective \cite[5.4]{Ki}, but the hyperk\"ahler case
1033: remains open.  Thus Theorems \ref{htm} and \ref{hm} can be interpreted as saying
1034: that the Kirwan maps for hypertoric varieties are surjective, and computing the kernel.
1035: Likewise, Theorems \ref{htsm} and \ref{hsm} assert that the $S^1$-equivariant Kirwan maps
1036: $$\ktdso:H^*_{\Tn\times S^1}(\cot)\to\htsm$$ and
1037: $$\kso:H^*_{\Tk\times S^1}(\cot)\to\hsm$$ are surjective,
1038: and provide computations of their kernels.
1039: \end{remark}
1040: 
1041: In order to apply Konno's results, we will make use of
1042: the principle of {\em equivariant formality},
1043: proven for compact manifolds in \cite{Ki}, which we adapt to our situation
1044: in Proposition \ref{formality}.
1045: For the sake of simplicity, we will restrict our attention
1046: to the case where $X$ is compact and nonempty.
1047: This condition will be necessary for the application of
1048: Proposition \ref{formality} and the proof of Theorem \ref{tara},
1049: both of which require a proper Morse function, which we get from
1050: Proposition \ref{proper}.
1051: We note, however, that both Proposition \ref{formality} and Theorem
1052: \ref{tara} can be extended to the case of a general hypertoric
1053: variety by a Mayer-Vietoris argument, using the fact that the core
1054: $C\subs M$ is a compact $\Td\times S^1$-equivariant deformation retract.
1055: We present the slightly less general Morse theoretic proofs
1056: only because we find them more pleasant.
1057: 
1058: \begin{proposition}\label{formality}
1059: Let $M$ be a symplectic
1060: orbifold, possibly noncompact but of finite topological type.
1061: Suppose that $M$ admits a hamiltonian action of a torus $T\times S^1$,
1062: and that the $S^1$-component $\Phi:M\to\R$ of the moment map
1063: is proper and bounded below.  Then 
1064: $H^*_{T\times S^1}(M)$ 
1065: is a free module over $H^*_{S^1}(pt)$.
1066: \end{proposition} 
1067: 
1068: \begin{proof}
1069: Because $\Phi$ is a moment map, it is a Morse-Bott function
1070: such that all of the critical suborbifold and their normal bundles
1071: carry almost complex structures.  Thus we get a Morse-Bott stratification
1072: of $M$ into even-dimensional $T$-invariant suborbifolds.
1073: This tells us, as in [Ki, 5.8],
1074: that the spectral sequence associated to the fibration
1075: $M\hookto EG\times_G M\to BG$ collapses, and we get the desired result.
1076: \end{proof}
1077: 
1078: Consider the following commuting square of maps, where
1079: $\phi$ and $\psi$ are each given by setting $x$ to zero.
1080: 
1081: $$\begin{CD}
1082: H^*_{\Tn\times S^1}(\cot) @>\ktdso >> \htsm\\
1083: @V\phi VV @ VV\psi V\\
1084: H^*_{\Tn}(\cot) @>\ktd >> \htm\\
1085: \end{CD}$$
1086: 
1087: Proposition \ref{formality} has the following consequence.
1088: 
1089: \begin{corollary}\label{enough}
1090: Let $\mathcal{I}\subs\ker\ktdso$ be an ideal with $\phi(\mathcal{I}) = \ker\ktd$.
1091: Then $\mathcal{I}=\ker\ktdso$.
1092: \end{corollary}
1093: 
1094: \begin{proof}
1095: Suppose that $a\in\ker\ktdso\smallsetminus\mathcal{I}$ is a homogeneous class of minimal degree,
1096: and choose $b\in\mathcal{I}$ such that $\phi(a-b)=0$.  Then $a-b = cx$ for some
1097: $c\in H^*_{\Tn\times S^1}(\cot)$.  By Proposition \ref{formality},
1098: $cx\in\ker\ktdso\impl c\in\ker\ktdso$, hence $c\in\ker\ktdso\smallsetminus\mathcal{I}$
1099: is a class of lower degree than $a$.
1100: \end{proof}
1101: 
1102: \vspace{-\baselineskip}
1103: \begin{lemma}\label{surjective}
1104: The equivariant Kirwan map $\ktdso$ is surjective.
1105: \end{lemma}
1106: 
1107: \begin{proof}
1108: Suppose that $\gamma\in\htsm$ is a homogeneous class of minimal degree
1109: that is {\em not}
1110: in the image of $\ktdso$.  By Theorem \ref{htm} $\ktd$ is surjective, hence we may choose a class
1111: $\eta\in\phi^{-1}\ktd^{-1}\psi(\gamma)$.
1112: Then $\ktdso(\eta)-\gamma = x\delta$ for some $\delta\in\htsm$,
1113: and therefore $\delta$ is a class of lower degree that is not in the image of $\ktdso$.
1114: \end{proof}
1115: 
1116: \vspace{-\baselineskip}
1117: \begin{proofhtsm}
1118: For any element $h\in H^2_{T^n\times S^1}(\Hn;\Z)$,
1119: let $\tilde{L}_h = \Hn\times\C_{h}$ 
1120: be the $\Tn\times S^1$-equivariant line bundle on $\Hn$
1121: with equivariant Euler class $h$. 
1122: This gives 
1123: $\tilde{L}_h$, as well as its dual $\tilde{L}_h^*$, 
1124: the structure of a \(T^n \times S^1\)-equivariant bundle. 
1125: Let $$L_h = \tilde{L}_h|_{\bar{\mu}_{\C}^{-1}(0)^{\text{ss}}}/T_{\C}^k$$ 
1126: be the quotient $\Td\times S^1$-equivariant line bundle on $M$.
1127: Let $\{u_i\}$ be the standard basis of $\Tnd$.
1128: Identifying $H^2_{T^n\times S^1}(\Hn;\Z)$ with $\Tnd\oplus\Z x$,
1129: We will use $\tilde L_i$ to denote the bundle $\tilde L_{u_i\oplus 0}$, 
1130: and $\tilde K$ to denote the bundle $\tilde L_{0\oplus x}$,
1131: with quotients $L_i$ and $K$. 
1132: Since the $\Tdso$-equivariant Euler class $e(L_i)$
1133: is the image of $u_i\oplus 0$ under the hyperk\"ahler Kirwan map
1134: $H^*_{T^n\times S^1}(\Hn)\to \htsm$, we will abuse notation and 
1135: denote it by $u_i$.  Similarly, we will denote $e(K)$ by $x$.
1136: Corollary \ref{surjective} tells us that
1137: $\htsm$ is generated by $u_1,\ldots,u_n,x$.
1138: 
1139: Consider the $\Tn\times S^1$-equivariant section $\tilde s_i$ of $\tilde{L}_i$
1140: given by the function $\tilde s_i(z,w)=z_i$.
1141: This descends to a $\Tdso$-equivariant section $s_i$ of $L_i$
1142: with zero-set $Z_i
1143: := \{[z,w]\in M\mid z_i=0\}$. 
1144: Similarly, the function $\tilde t_i(z,w)=w_i$ defines a $\Tdso$-equivariant
1145: section of $L_i^*\otimes K$ with zero set $W_i := \{[z,w]\in M\mid w_i=0\}$.
1146: Thus the divisor $Z_i$ represents the cohomology class $u_i$,
1147: and $W_i$ represents $x-u_i$.
1148: Note, by the proof of Lemma \ref{sides},
1149: that $\mr(Z_i) = G_i$ and $\mr(W_i) = F_i$ for all $1\leq i \leq n$.
1150: 
1151: Let $S = S_1\sqcup S_2$ be a subset of $\{1,\ldots n\}$ such that
1152: $\big(\cap_{i\in S_1}G_i\big)\cap
1153: \big(\cap_{j\in S_2}F_j\big) = \emptyset,$ and hence
1154: $$\big(\cap_{i\in S_1}Z_i\big)\cap
1155: \big(\cap_{j\in S_2}W_j\big)\subs\mr^{-1}\bigg(
1156: \big(\cap_{i\in S_1}G_i\big)\cap
1157: \big(\cap_{j\in S_2}F_j\big)\bigg) = \emptyset.$$
1158: Now consider the vector bundle
1159: $E_S=\left(\oplus_{i\in S_1}L_i\right)\oplus\left(\oplus_{j\in S_2}L_j^*\otimes K\right)$
1160: with equivariant Euler class $$e(E_S)=\prod_{i\in S_1}u_i\times\prod_{j\in S_2}(x-u_j).$$ 
1161: The section $\left(\oplus_{i\in S_1}s_i\right)\oplus\left(\oplus_{i\in S_2}t_i\right)$ 
1162: is a nonvanishing equivariant
1163: global section of $E_S$,
1164: hence for any such $S$, $e(E_S)$ is trivial in $\htsm$.
1165: 
1166: The fact that $u_1,\ldots,u_n,x$ generate $\htsm$ is proven in Lemma \ref{surjective},
1167: and the fact that
1168: we have found all of the relations
1169: follows from Theorem \ref{htm} and Corollary \ref{enough}.
1170: \end{proofhtsm}
1171: 
1172: \vspace{-\baselineskip}
1173: \begin{proofhsm}
1174: The proof of this theorem is identical to the proof of Theorem \ref{htsm},
1175: making use of Theorem \ref{hm} rather than Theorem \ref{htm}.
1176: \end{proofhsm}
1177: 
1178: How sensitive are the invariants $\htsm$ and $\hsm$?
1179: % By Proposition \ref{formality}, 
1180: We can recover $\htm$ and $\hm$
1181: by setting $x$ to zero, hence they are at least as fine
1182: as the ordinary or $\Td$-equivariant cohomology rings.
1183: The ring $\htsm$ does {\em not} depend on coorientations,
1184: for if $M'$ is related to $M$ by flipping the coorientation
1185: of the $l^{\text{th}}$ hyperplane $H_k$, then the map taking $u_i$ to
1186: $u_i$ for $i\neq l$ and $u_l$ to $x-u_l$ is an isomorphism
1187: between $\htsm$ and $H^*_{\Td\times S^1}(M')$.
1188: It is, however, dependent on the affine structure of 
1189: the arrangement $\arr$.
1190: 
1191: \begin{example}\label{ts}
1192: We compute the equivariant cohomology ring $\htsm$
1193: for the hypertoric varieties $M_a$, $M_b$, and $M_c$
1194: defined by the arrangements in Figure 2(a), (b), and (c), respectively.
1195: $$H^*_{\Td\times S^1}(M_a)=\Z[u_1,\ldots,u_4,x]\big/
1196: \< u_2u_3, u_1(x-u_2)u_4, u_1u_3u_4 \>,$$
1197: $$H^*_{\Td\times S^1}(M_b)=\Z[u_1,\ldots,u_4,x]\big/
1198: \< (x-u_2)u_3, u_1u_2u_4, u_1u_3u_4 \>,$$
1199: $$H^*_{\Td\times S^1}(M_c)=\Z[u_1,\ldots,u_4,x]\big/
1200: \< u_2u_3, (x-u_1)u_2(x-u_4), u_1u_3u_4 \>.$$
1201: As we have already observed, $H^*_{\Td\times S^1}(M_a)$ and $H^*_{\Td\times S^1}(M_b)$
1202: are isomorphic by interchanging $u_2$ with $x-u_2$.
1203: % We now show that $H^*_{\Td\times S^1}(M_a)$ is not isomorphic
1204: % to $H^*_{\Td\times S^1}(M_c)$.  
1205: One can check that the annihilator of $u_2$ in $H^*_{\Td\times S^1}(M_a)$
1206: is the principal ideal generated by $u_3$, while the ring $H^*_{\Td\times S^1}(M_c)$
1207: has no degree $2$ element whose annihilator is generated by a single element of degree $2$.
1208: Hence $H^*_{\Td\times S^1}(M_c)$ is not isomorphic to the other two rings.
1209: % Since $(x-u_2)(x-u_3)$ is the unique
1210: % degree $4$ relation in both rings, any isomorphism 
1211: % between the two rings
1212: % would have to have to take $x-u_2$ and $x-u_3$ 
1213: % to scalar multiples of $x-u_2$ and $x-u_3$, possibly in the reverse order.
1214: % One can check that the only degree $4$ classes
1215: % in $H^*_{\Td\times S^1}(M_a)$ that annihilate $x-u_2$ are multiples of $x-u_3$,
1216: % whereas in $H^*_{\Td\times S^1}(M_c)$, $x-u_2$ is killed by $u_1u_4$
1217: % and $x-u_3$ is killed by $(x-u_1)(x-u_4)$.
1218: % Hence $x-u_2$ can be taken neither to a multiple of $x-u_2$ nor of $x-u_3$,
1219: % therefore the rings cannot be isomorphic.
1220: \end{example}
1221: 
1222: The ring $\hsm$, on the other hand, is sensitive to coorientations
1223: as well as the affine structure of $\arr$.
1224: 
1225: \begin{example}\label{s}
1226: We now compute the ring $\hsm$ for 
1227: $M_a$, $M_b$, and $M_c$ of Figure 2.
1228: Theorem \ref{hsm} tells us that we need only to quotient the ring
1229: $\htsm$ by $\ker(\iota^*)$.
1230: For $M_a$, the kernel of $\iota_a^*$ is generated by
1231: $u_1+u_2-u_3$ and $u_1-u_4$, hence we have
1232: \begin{eqnarray*}
1233: H^*_{S^1}(M_a)&=&\Z[u_2,u_3,x]\big/
1234: \< u_2u_3, (u_3-u_2)^2(x-u_2), (u_3-u_2)^2u_3 \>\\
1235: &\cong & \Z[u_2,u_3,x]\big/\< u_2u_3, (u_3-u_2)^2(x-u_2), u_3^3 \>.\\
1236: \end{eqnarray*}
1237: Since the hyperplanes of 2(c) have the same coorientations as those
1238: of 2(a), we have $\ker\iota_b^*=\ker\iota_a^*$, hence
1239: \begin{eqnarray*}
1240: H^*_{S^1}(M_c) &=& \Z[u_2,u_3,x]\big/
1241: \< u_2u_3, (x-u_3+u_2)^2u_2, (u_3-u_2)^2u_3 \>\\
1242: &\cong & \Z[u_2,u_3,x]\big/\< u_2u_3, (x-u_3+u_2)^2u_2, u_3^3 \>.\\
1243: \end{eqnarray*}
1244: Finally, since Figure 2(b) is obtained from 2(a) by flipping the coorientation
1245: of $H_2$, we find that $\ker(\iota_b^*)$ is generated by $u_1-u_2-u_3$
1246: and $u_1-u_4$, therefore
1247: $$H^*_{S^1}(M_b)=\Z[u_2,u_3,x]\big/
1248: \< (x-u_2)u_3, (u_2+u_3)^2u_2, (u_2+u_3)^2u_3 \>.$$
1249: As in Example \ref{ts}, $H^*_{S^1}(M_a)$ and $H^*_{S^1}(M_c)$
1250: can be distinguished by the fact that the annihilator
1251: of $u_2\in H^*_{S^1}(M_a)$ is generated by a single element of degree $2$,
1252: and no element of $H^*_{S^1}(M_c)$ has this property.
1253: On the other hand, $H^*_{S^1}(M_b)$ is distinguished from 
1254: $H^*_{S^1}(M_a)$ and $H^*_{S^1}(M_c)$ by the fact that neither $x-u_2$ nor $u_3$ cubes to zero.
1255: % Each of three rings have two linear elements, unique up to scale, that multiply to zero.
1256: % In $H^*_{S^1}(M_c)$, both of them cube to zero, while
1257: % in $H^*_{S^1}(M_a)$ and $H^*_{S^1}(M_b)$, only $(x-u_3)$ cubes to zero.
1258: % By changing coordinates, we can write
1259: % $$H^*_{S^1}(M_a) = \Z[x,y,z]\big/\<yz, (z-y+x)^2(x-y), z^3\>$$ and
1260: % $$H^*_{S^1}(M_b) = \Z[x,y,z]\big/\<yz, (z-y)^2(x-y), z^3\>.$$
1261: % Any isomorphism from $H^*_{S^1}(M_a)$ to $H^*_{S^1}(M_b)$ would have to take
1262: % $y$ and $z$ to multiples of themselves, and $x$ to an expression of the form
1263: % $\a x+\b y+ \gamma z$, with $\a\neq 0$.  Then a contradiction arises from the
1264: % fact that the first degree six relation in $H^*_{S^1}(M_a)$ is cubic in $x$,
1265: % while $H^*_{S^1}(M_b)$ has no such nontrivial relations.
1266: % Hence all three rings are nonisomorphic.
1267: \end{example}
1268: 
1269: \begin{remark}\label{lawrence}
1270: Theorems \ref{htsm} and \ref{hsm} can be interpreted
1271: in light of the recent work of Hausel and Sturmfels \cite{HS} 
1272: on Lawrence toric varieties.
1273: The Lawrence toric variety $N$ associated to the arrangement $\arr$
1274: is the K\"ahler reduction $\cot\mod\Tk$,
1275: so that $M$ sits inside of $N$ as the complete
1276: intersection cut out by the equation $\bar\mc(z,w)=0$.
1277: The residual torus acting on $N$ has dimension $d+n$,
1278: and includes the $(d+1)$-dimensional torus $\Td\times S^1$
1279: acting on $M$, and
1280: the inclusion of $M$ into
1281: $N$ induces an isomorphism on $\Td\times S^1$-equivariant cohomology.
1282: One can use geometric arguments similar to those that were applied to prove
1283: Theorem \ref{htsm},
1284: or the purely combinatorial approach of \cite{HS},
1285: to show that
1286: $$H_{T^{d+n}}^*(N)=\Q[u_1,\ldots,u_n,v_1,\ldots,v_n]\bigg/
1287: \< \prod_{i\in S_1}u_i\times\prod_{j\in S_2}v_j\hs\bigg{\vert}\hs\bigcap_{i\in S} H_i = \emptyset\>.$$
1288: From here we can recover $\htsm = H_{\Td\times S^1}^*(N)$
1289: by setting $u_i+v_i = u_j+v_j$ for all $i,j\leq n$.
1290: Note that Hausel and Sturmfels' work applies to the general orbifold case.
1291: \end{remark}
1292: \end{section}
1293: 
1294: \begin{section}{A deformation of the Orlik-Solomon algebra of \boldmath$\arr$}\label{defos}
1295: Let $M_{\R} \subs M$ be the {\em real locus}
1296: $\{[z,w]\in M\mid z,w \text{ real}\}$ of $M$ with respect to the complex structure $J_1$.
1297: The full group $\Td\times S^1$ does not act on $M_{\R} $,
1298: but the subgroup $\Tdr\times\Zt$ does act,
1299: where $\Tdr :=\Z_2^d\subs \Td$ is the fixed point set
1300: of the involution of $\Td$ given by complex
1301: conjugation.\footnote{It is interesting to note that
1302: the real locus with respect to the complex structure $J_1$
1303: is in fact a complex submanifold with respect to the
1304: one of the other complex structures on $M$.  
1305: The action of $\Tdr$ is holomorphic because $\Tdr$ is a subgroup of $\Td$,
1306: which preserves all of the complex structures on $M$.
1307: The action of $\Zt$, on the other hand, is anti-holomorphic,
1308: i.e. it can be thought of as complex conjugation.}
1309: In this section we will study the geometry of the real locus,
1310: focusing in particular on the properties of the residual $\Zt$ action.
1311: 
1312: A proof of a more general statement of the following theorem is 
1313: forthcoming in \cite{HH}. 
1314: 
1315: % BEGIN NEW
1316: % We begin by 
1317: % extending a theorem of
1318: % \cite{BGH, Sc}.
1319: % stating a theorem about the equivariant cohomology of $M_{\R} $,
1320: % the proof of which is forthcoming in \cite{HH}.
1321: 
1322: \begin{theorem}\label{tara}
1323: Let $G = \Tdso$ or $\Td$,
1324: and $G_{\R} = \Tdr\times\Zt$ or $\Tdr$.
1325: Then we have $H^{2*}_{G}(M;\Zt)\cong
1326: H^{*}_{G_{\R}}(M_{\R} ;\Zt),$
1327: i.e. the rings are isomorphic
1328: by an isomorphism that halves the grading.
1329: Furthermore, this isomorphism identifies the class 
1330: $u_i\in H^{*}_{G}(M;\Zt)$, 
1331: represented by the divisor $Z_i$, with the class in $H^*_{G_{\R}}(M_{\R} ;\Zt)$
1332: represented by the divisor $Z_i\cap M_{\R} $, and likewise takes
1333: $x-u_i$ (if $G=\Tdso$) or
1334: $-u_i$ (if $G=\Td$) to the class represented by $W_i\cap M_{\R} $. 
1335: \end{theorem}
1336: 
1337: \begin{sketch}
1338: Consider the injection $H^*_G(M;\Zt)\hookto H^*_G(M^G;\Zt)$
1339: given by the inclusion of the fixed point set into $M$.
1340: % Let \(M^{(1)}\) be the {\em one-skeleton} of $M$, i.e. the set of 
1341: % points in $M$ whose $G$-orbits are 0- or 1-dimensional. 
1342: % Let \(M^{(0)}\) denote the fixed points. Let 
1343: % \(j: M^{(1)} \hookrightarrow M\) and \(i: M^{(0)} \hookrightarrow M\) 
1344: % denote the natural inclusions. 
1345: The essential idea is to show that 
1346: % the image \(j^{*}(H^*_{G}(M^{(1)})
1347: % \subset H^{*}_{G}(M^{(0)})\) is the same as the image 
1348: a class in $H^*_G(M^G;\Zt)$ extends over $M$ if and only
1349: if it extends to the set of points on which $G$ acts with
1350: a stabilizer of codimension at most $1$,
1351: and then to show that a similar
1352: statement in $G_{\R}$-equivariant cohomology 
1353: also holds for the real locus $M_{\R} $ with its $G_{\R}$
1354: action. One then uses a canonical isomorphism \(H^2_G(pt,\Z_2) \cong 
1355: H^1_{G_{\R}}(pt,\Z_2)\) to give the result. 
1356: 
1357: The key to the proof is a noncompact $G_{\R}$ version
1358: of the proposition in 
1359: \cite{TW} stating that the $G_{\R}$-equivariant Euler class of 
1360: the negative normal bundle of a critical point $p$ 
1361: is not a zero divisor, which can be shown explicitly
1362: using a local normal form for the actions of $G$ and $G_{\R}$. 
1363: % Using a local normal form theorem for 
1364: % Hamiltonian $G$-manifolds equipped with a compatible 
1365: % antisymplectic involution $\sigma$ and properties of 
1366: % the $G$ weights and $G_{\R}$ weights at the fixed points, 
1367: % one can compute
1368: % explicitly the $G_{\R}$ equivariant Euler class and verify 
1369: %that it is not a zero divisor.  
1370: The proposition then follows 
1371: from standard $G_{\R}$ versions of the Thom isomorphism theorem
1372: with coefficients in $\Z_2$. 
1373: Since a component of the 
1374: moment map is proper, bounded below, and has finitely many
1375: fixed points, one can then check that the inductive argument, given 
1376: in Section 3 of \cite{TW}
1377: to complete the proof of \cite[Thm 1]{TW} also holds in this case.
1378: \end{sketch}
1379: 
1380: % What kind of smoothness assumptions do we need to make about 
1381: % $\arr$ and $M$?
1382: % In this section we work with cohomology with coefficients in $\Zt$,
1383: % therefore it is okay for $M$ to have orbifold points
1384: % as long as they are of odd order.  In terms of $\arr$, this means
1385: % that the primitive normal vectors to the hyperplanes at each vertex must
1386: % span a sublattice of odd index inside of $\liet^d_{\Z}$.
1387: % Such an arrangement will be referred to as an {\em odd} arrangement.
1388: % The oddness condition may seem only marginally more general than
1389: % the smoothness condition, but it is significant because of the following
1390: % important lemma:
1391: % 
1392: % \begin{lemma}\label{odd}
1393: % Every simple arrangement is combinatorially equivalent 
1394: % as an affine arrangement to
1395: % an odd arrangement.
1396: % \end{lemma}
1397: % 
1398: % \begin{proof}
1399: % \end{proof}
1400: % 
1401: % \noindent For the remainder of this section, we will assume that $\arr$ is odd.
1402: 
1403: Let us consider the restriction of the hyperk\"ahler moment map
1404: $\mhk=\mr\oplus\mc$ to $M_{\R} $.  Since $z$ and $w$ are real
1405: for every $[z,w]\in M_{\R} $, the map $\mc$
1406: takes values in $\td_{\R}\subs\td_{\C}$, which we will identify
1407: with $i\R^n$, so that $f = \mr|_{M_{\R}} \oplus\mc|_{M_{\R}}$ takes values
1408: in $\R^n\oplus i\R^n\cong\C^n$.
1409: Note that $f$ is $\Zt$-equivariant, with $\Zt$ acting on
1410: $\Cn$ by complex conjugation.
1411: 
1412: \begin{lemma}
1413: The map $f:{M_{\R}}\to\Cn$ is surjective, and the fibers are the orbits of $\Tdr$.
1414: The stabilizer of a point $x\in {M_{\R}}$ has order 
1415: $2^r$, where $r$ is the number of hyperplanes in the complexified
1416: arrangement $\arr_{\C}$ containing the point $f(x)$.
1417: \end{lemma}
1418: 
1419: \begin{proof}
1420: For any point $p=a+bi\in\Cn$, choose a point $[z,w]\in M$ such that
1421: $\mr[z,w]=a$ and $\mc[z,w]=b$.  We can hit $[z,w]$
1422: with an element of $\Td=\Tn/\Tk$ to make $z$ real, and the fact that
1423: $\mc[z,w]\in\R^n$ forces $w$ to be real as well, hence we may assume
1424: that $[z,w]\in {M_{\R}}$.
1425: Then $f^{-1}(p)=\mu_{\text{HK}}^{-1}\left(a,b\right)\cap {M_{\R}}
1426: = \Td[z,w]\cap {M_{\R}} = \Tdr[z,w]$.
1427: The second statement follows easily from \cite[3.1]{BD}.
1428: \end{proof}
1429: 
1430: Let $Y\subs {M_{\R}}$ be the locus of points on which $\Tdr$
1431: acts freely, i.e. the preimage under $f$ of the space 
1432: $\comp:=\bomp$.
1433: The inclusion map $Y\hookto {M_{\R}}$
1434: induces maps backward on cohomology, which we will denote
1435: $$\phi:\hr\to H^*_{\Tdr}(Y;\Zt)\cong H^*(\comp;\Zt)$$
1436: and
1437: $$\phi_2:\hrs\to H^*_{\Tdr\times\Zt}(Y;\Zt)\cong H^*_{\Zt}(\comp;\Zt).$$
1438: The ring $H^*(\comp;\Z)$ has been studied extensively, and is called
1439: the {\em Orlik-Solomon algebra} \cite{OT}, which we will denote
1440: by $\os$.
1441: A remarkable fact about the Orlik-Solomon algebra is that it depends
1442: only on the combinatorial structure of $\arr$; the following is a presentation
1443: in terms of {\it anticommuting} generators $e_1,\ldots,e_n$ \cite{OS}:
1444: $$\os\cong H^*(\comp;\Z)\cong\Z[e_1,\ldots,e_n]\big/
1445: \< \Pi_{i\in S}e_i\mid\cap_{i\in S} H_i = \emptyset \>.$$
1446: Rather than working with anticommuting generators, we can work over the ground field $\Zt$,
1447: in which case commutativity and anticommutativity are the same.
1448: Because $\os$ is torsion-free \cite[3.74]{OT}, we have
1449: $$\ost\cong H^*(\comp;\Zt)\cong\Zt[e_1,\ldots,e_n]\big/
1450: \< \Pi_{i\in S}e_i\mid\cap_{i\in S} H_i = \emptyset \>
1451: +\< e_i^2\mid i\leq n\>,$$
1452: where $\operatorname{deg}(e_i)=1$.
1453: 
1454: \begin{claim}
1455: The map $\phi:H^*_{\Tdr}(M_{\R};\Zt)\to\ost$ 
1456: takes $u_i$ to $e_i$, hence $\ker\phi$
1457: is generated by
1458: the set $\{u_i^2\mid i\leq n\}$.
1459: \end{claim}
1460: 
1461: \begin{proof}
1462: Recall from Section \ref{ht}
1463: that the hyperplane $H_i\subs \tdd$
1464: is defined by the equation $\< x,a_i\> = \< -\tilde\a,\e_i\>$.
1465: Let $\eta_i:\Cn\to\C$ be the affine map
1466: taking $x$ to $\< x,a_i\> + \< \tilde\a,\e_i\>$,
1467: so that $H_i^{\C}$ is cut out of $\Cn$ by $\eta_i$.
1468: Then $\eta_i$ restricts to a map $\comp\to\C^*$, and
1469: Orlik and Terao identify $e_i$ with the cohomology class
1470: represented by the pull-back
1471: of the submanifold $\R_-$ (the negative reals) along $\eta_i$ [OT, 5.90].
1472: Theorem \ref{tara} tells us that the cohomology
1473: class $u_i$ is represented by the divisor $Z_i\cap {M_{\R}}$.
1474: By Lemma \ref{sides}, $f(Z_i\cap {M_{\R}}) = G_i\cap \R^n$,
1475: hence $\phi(u_i)$ is the class represented by the submanifold
1476: $G_i\cap \R^n\cap \comp = 
1477: \eta_i^{-1}(\R_-)$.
1478: \end{proof}
1479: 
1480: \vspace{-\baselineskip}
1481: \begin{remark}\label{kernel}
1482: The fact that the classes $u_i^2$ lie in the kernel of $\phi$ can be seen
1483: by noting that $u_i$ is represented
1484: both by the divisor $Z_i = \{[z,w]\in {M_{\R}}\mid z_i=0\}$, 
1485: and, since $u_i = -u_i$ over $\Zt$, 
1486: also by the divisor $W_i = \{[z,w]\in {M_{\R}}\mid w_i=0\}$.
1487: The condition $x\in Z_i\cap W_i$ says exactly that
1488: $\mr(x)\in H_i$, therefore
1489: the intersection $Z_i\cap W_i \cap Y$ is empty.
1490: \end{remark}
1491: 
1492: In some sense we have cheated here; we have concluded that
1493: we can recover a presentation of $\ost$ from a presentation
1494: of $\htm$, but we used the fact that we already have a presentation
1495: of $\ost$.  In the $\Zt$-equivariant picture, however,
1496: our trivial observation turns magically into new information,
1497: giving us a presentation of the equivariant cohomology
1498: ring $\hscomp$.
1499: 
1500: \begin{theorem}\label{surj}
1501: The map $\phi_2$ is surjective, with kernel generated by 
1502: $\{u_i (x-u_i)\mid i\leq n\}$.
1503: Hence 
1504: $$\hscomp \cong H^*_{\Tdr\times\Zt}(M_{\R};\Zt)\big/\ker\phi_2$$
1505: $$\cong\Zt[u_1,\ldots,u_n,x]\big/
1506: \<\Pi_{i\in S_1}u_i\times\Pi_{j\in S_2}(x-u_j)
1507: \mid\cap_{i\in S} H_i = \emptyset\>
1508: +\<u_i(x-u_i)\mid i\leq n\>.$$
1509: \end{theorem}
1510: 
1511: \begin{proof}
1512: % Theorem \ref{tara} and Proposition \ref{formality}
1513: % imply that $$\hrs\cong\hr\otimes H_{\Zt}^*(pt)$$
1514: % as a module over $H_{\Zt}^*(pt)$.
1515: % On the other hand, 
1516: The fact that $\arr_{\C}$ is the 
1517: complexification of a real arrangement
1518: tells us that the linear forms $\eta_i$ have real coefficients,
1519: therefore the generators $e_i$ of $\hcomp$ are $\Zt$-invariant.
1520: This implies, by \cite[3.5]{Bo},
1521: that $\hscomp$
1522: is a free module over $H_{\Zt}^*(pt)$.
1523: Then surjectivity of $\phi_2$ follows from surjectivity of $\phi$
1524: using a formal argument identical to that of Corollary \ref{surjective}.
1525: By Theorem \ref{tara} and Proposition \ref{formality},
1526: $\hrs$ is a free module over $H_{\Zt}^*(pt)$,
1527: therefore $\ker\phi_2$ is a free 
1528: $H_{\Zt}^*(pt)$-module of rank $n$.
1529: The fact that $u_i(x-u_i)\in \ker\phi_2$ follows from the
1530: argument of Remark \ref{kernel}, hence we are done.
1531: \end{proof}
1532: 
1533: The ring $\hscomp$ is therefore a deformation
1534: of $\ost$ (over the base $\operatorname{Spec}\Zt[x]$)
1535: that depends nontrivially on the affine structure of $\arr$,
1536: rather than simply on the underlying matroid.
1537: 
1538: \begin{example}\label{first}
1539: Consider the arrangements $\arr_a$ and $\arr_c$ in Figure 2(a) and 2(c).
1540: By Theorem \ref{surj} and Example \ref{ts} we have
1541: $$H^*_{\Zt}(\mathcal{M}(\arr_a);\Zt)\cong
1542: \Zt [u_1,\ldots,u_4,x]\bigg/
1543: \< \begin{array}{c}
1544: u_1(x-u_1), u_2(x-u_2), u_3(x-u_3), u_4(x-u_4),\\
1545: u_2u_3, u_1(x-u_2)u_4, u_1u_3u_4 
1546: \end{array}\>$$
1547: and
1548: $$H^*_{\Zt}(\mathcal{M}(\arr_c);\Zt)\cong
1549: \Zt [u_1,\ldots,u_4,x]\bigg/
1550: \< \begin{array}{c}
1551: u_1(x-u_1), u_2(x-u_2), u_3(x-u_3), u_4(x-u_4),\\
1552: u_2u_3, (x-u_1)u_2(x-u_4), u_1u_3u_4
1553: \end{array}\>.$$
1554: The map $f:H^*_{\Zt}(\mathcal{M}(\arr_a);\Zt)\to H^*_{\Zt}(\mathcal{M}(\arr_b);\Zt)$
1555: given by
1556: $$f(u_1) = u_1+u_2, f(u_2) = u_2+u_3+x, f(u_3) = u_3,
1557: f(u_4) = u_2+u_4,\text{ and }f(x)=x$$
1558: is an isomorphism of graded $\Zt[x]$-algebras, hence the
1559: ring $\hscomp$ is not a complete invariant of smooth, rational, 
1560: affine arrangements up to combinatorial equivalence.\footnote{We thank
1561: Graham Denham for finding this isomorphism.}
1562: \end{example}
1563: 
1564: \begin{example}\label{second}
1565: Now consider the arrangements $\arr_a'$ and $\arr_c'$ obtained from $\arr_a$ and $\arr_c$
1566: by adding a vertical line on the far left, as shown below.
1567: \begin{figure}[h]
1568: \begin{center}
1569: \psfrag{Ha'}{$\arr_a'$}
1570: \psfrag{Hc'}{$\arr_c'$}
1571: \includegraphics{htv3.eps}
1572: \end{center}
1573: \end{figure}
1574: Again by Theorem \ref{surj}, we have
1575: $$H^*_{\Zt}(\mathcal{M}(\arr_a');\Zt)\cong
1576: \Zt [\vec{u},x]\bigg/
1577: \< \begin{array}{c}
1578: u_1(x-u_1), u_2(x-u_2), u_3(x-u_3), u_4(x-u_4),\\ u_5(x-u_5), 
1579: u_2u_3, (x-u_1)u_5, u_1(x-u_2)u_4,\\ u_1u_3u_4, (x-u_2)u_4u_5, u_3u_4u_5 
1580: \end{array}\>$$
1581: and
1582: $$H^*_{\Zt}(\mathcal{M}(\arr_c');\Zt)\cong
1583: \Zt [\vec{u},x]\bigg/
1584: \< \begin{array}{c}
1585: u_1(x-u_1), u_2(x-u_2), u_3(x-u_3), u_4(x-u_4),\\ u_5(x-u_5),
1586: u_2u_3, (x-u_1)u_5, (x-u_1)u_2(x-u_4),\\ u_1u_3u_4, (x-u_2)u_4u_5, u_3u_4u_5 
1587: \end{array}\>.$$
1588: We have used Macaulay 2 to check that the annihilator of the element 
1589: $u_2\in H^*_{\Zt}(\mathcal{M}(\arr_a');\Zt)$ is generated by two linear elements
1590: (namely $u_3$ and $x-u_2$) and nothing else, while there is no element
1591: of $H^*_{\Zt}(\mathcal{M}(\arr_c');\Zt)$ with this property.
1592: Hence the two rings are not isomorphic.
1593: \end{example}
1594: \end{section}
1595: 
1596: \footnotesize{
1597: \begin{thebibliography}{10}
1598: 
1599: \bibitem[BD]{BD}
1600: R.~Bielawski and A.~Dancer.
1601: \newblock The geometry and topology of toric hyperk\"ahler manifolds.
1602: \newblock {\em Comm. Anal. Geom.} 8 (2000), 727--760.
1603: 
1604: \bibitem[BGH]{BGH}
1605: D.~Biss, T.~Holm, and V.~Guillemin.
1606: \newblock The mod 2 cohomology of fixed point sets
1607: of anti-symplectic involutions.
1608: \newblock math.SG/0107151.
1609: 
1610: \bibitem[Bo]{Bo}
1611: A.~Borel.
1612: \newblock {\em Seminar on Transformation Groups}, chapter XII.
1613: \newblock Annals of Mathematical Studies 46,
1614: \newblock Princeton, 1960.
1615: 
1616: \bibitem[BJ]{BJ}
1617: T.~Brocker and K.~Janich.
1618: \newblock {\em Introduction to Differential Topology}.
1619: \newblock Cambridge University Press, 1982.
1620: 
1621: % \bibitem[Br]{Br}
1622: % R.~Bryant.
1623: % \newblock An introduction to Lie groups and symplectic geometry.
1624: % \newblock {\em Geometry and quantum field theory (Park City, UT, 1991)},
1625: % 5--181.
1626: % 
1627: \bibitem[HH]{HH}
1628: M.~Harada and T.~Holm.
1629: \newblock The equivariant cohomology of hypertoric varieties and their real loci.
1630: \newblock In preparation.
1631: 
1632: \bibitem[HS]{HS}
1633: T.~Hausel and B.~Sturmfels.
1634: \newblock Toric hyperk\"ahler varieties.
1635: \newblock math.AG/0203096.
1636: 
1637: \bibitem[HL]{HL}
1638: P.~Heinzner and F.~Loose.
1639: \newblock Reduction of complex Hamiltonian $G$-spaces. 
1640: \newblock {\em Geom. Funct. Anal.} 4 (1994), no. 3, 288--297.
1641: 
1642: % \bibitem[Hi]{Hi}
1643: % N.~Hitchin.
1644: % \newblock Hyper-K\"ahler manifolds. 
1645: % \newblock S\'eminaire Bourbaki, Vol. 1991/92. 
1646: % \newblock {\em Astérisque} no. 206 (1992), Exp. no. 748, 3, 137--166.
1647: % 
1648: \bibitem[HKLR]{HKLR}
1649: N.~Hitchin, A.~Karlhede, U.~Lindstr\"om, and M.~Ro\v{c}ek.
1650: \newblock Hyper-K\"ahler metrics and supersymmetry.
1651: \newblock {\em Comm. Math. Phys.} 108 (1987)
1652: no. 4, 535--589.
1653: 
1654: \bibitem[Ki]{Ki}
1655: F.~Kirwan.
1656: \newblock {\em Cohomology of quotients in symplectic and algebraic geometry}.
1657: \newblock Mathematical Notes 31,
1658: \newblock Princeton University Press, 1984.
1659: 
1660: \bibitem[K1]{K1}
1661: H.~Konno.
1662: \newblock Cohomology rings of toric hyperk\"ahler manifolds.
1663: \newblock {\em Int. J. of Math.} 11 (2000), 
1664: no. 8, 1001--1026.
1665: 
1666: \bibitem[K2]{K2}
1667: H.~Konno.
1668: \newblock Equivariant cohomology rings of toric hyperk\"ahler manifolds.
1669: \newblock {\em Quaternionic structures in mathematics and physics (Rome, 1999)},
1670: 231--240 (electronic).
1671: 
1672: % \bibitem[MFK]{MFK}
1673: % D.~Mumford, J.~Fogarty, and F.~Kirwan.
1674: % \newblock {\em Geometric Invariant Theory}.
1675: % \newblock Third edition. Ergebnisse der Mathematik und ihrer 
1676: % Grenzgebiete (2) [Results in Mathematics and Related Areas (2)], 34.
1677: % \newblock Springer-Verlag, 1994.
1678: % 
1679: \bibitem[Na]{Na}
1680: H.~Nakajima.
1681: \newblock Instantons on ALE spaces, quiver varieties, and Kac-Moody algebras.
1682: \newblock {\em Duke Math. J.} 76 (1994) no. 2, 365--416.
1683: 
1684: \bibitem[OS]{OS}
1685: P.~Orlik and L.~Solomon.
1686: \newblock Combinatorics and topology of complements of hyperplanes.
1687: \newblock {\em Invent. Math.} 56 (1980), 167--189.
1688: 
1689: \bibitem[OT]{OT}
1690: P.~Orlik and H.~Terao.
1691: \newblock {\em Arrangements of Hyperplanes}.
1692: \newblock Grundlehren der mathematischen Wissenschaften 300,
1693: \newblock Springer-Verlag, 1992.
1694: 
1695: \bibitem[Pr]{Pr}
1696: N.~Proudfoot.
1697: \newblock {\em The equivariant Orlik-Solomon algebra}.
1698: \newblock http://math.berkeley.edu/$\sim$proudf/os.ps
1699: 
1700: \bibitem[Sc]{Sc}
1701: C.~Schmid.
1702: \newblock {\em Cohomologie \'equivariante des certaines vari\'et\'es
1703: hamiltoniennes et de leur partie r\'eelle.}
1704: \newblock Th\`ese \`a Universit\'e de Gen\`eve.
1705: \newblock Available at http://www.unige.ch/math/biblio/these/theses.html
1706: 
1707: \bibitem[Sj]{Sj}
1708: R.~Sjamaar.
1709: \newblock Holomorphic slices, symplecitic reduction, and multiplicities of representations.
1710: \newblock {\em Ann. Math.} 141 (1995), 87--129.
1711: 
1712: \bibitem[TW]{TW}
1713: S.~Tolman and J.~Weitsman.
1714: \newblock On the cohomology rings of Hamiltonian T-spaces.
1715: \newblock {\em Proc. of the Northern California Symplectic Geometry Seminar},
1716: Y. Eliashberg {\em et al.} eds. AMS Translations Series 2, vol. 196: 
1717: Advances in Mathematical Sciences 45 (1999), 251--258. 
1718: \end{thebibliography}
1719: }
1720: 
1721: \end{spacing}
1722: \end{document}
1723: