math0501017/ks3.tex
1: 
2: %preceeded by cat0.tex
3: \documentclass[12pt, reqno]{amsart}
4: \usepackage{graphicx}
5: %\usepackage[arrow,matrix,curve]{xy}
6: 
7: \newtheorem{theorem}{Theorem}[section]
8: \newtheorem{proposition}[theorem]{Proposition}
9: \newtheorem{lemma}[theorem]{Lemma}
10: \newtheorem{corollary}[theorem]{Corollary}
11: \newtheorem{conjecture}[theorem]{Conjecture}
12: 
13: 
14: \theoremstyle{definition} 
15: \newtheorem{definition}[theorem]{Definition}
16: \newtheorem{example}[theorem]{Example}
17: \newtheorem{remark}[theorem]{Remark}
18: \newtheorem{question}[theorem]{Question} 
19: 
20: 
21: \newcommand{\C}{\mathbb C} \newcommand{\N}{\mathbb N}
22: \newcommand{\R}{\mathbb R} \newcommand{\T}{\mathbb T}
23: \newcommand{\Z}{\mathbb Z}
24: 
25: \DeclareMathOperator{\dist}{{\rm dist}}
26: \DeclareMathOperator{\sys}{{\rm sys}}
27: \DeclareMathOperator{\stsys}{{\rm stsys}}
28: \DeclareMathOperator{\confsys}{{\rm confsys}}
29: \DeclareMathOperator{\pisys}{\sys\!\pi} 
30: 
31: \def\phisys {\phi{\rm sys}}
32: \DeclareMathOperator{\vol}{{\rm vol}}
33: 
34: \DeclareMathOperator{\area}{{\rm area}}
35: \DeclareMathOperator{\length}{{\rm length}}
36: 
37: \DeclareMathOperator{\Aut}{{\rm Aut}}
38: 
39: \DeclareMathOperator{\tauzero}{{\tau_0}}
40: \DeclareMathOperator{\FillVol}{{\rm FillVol}}
41: \DeclareMathOperator{\gmetric}{{\mathcal G}}
42: \DeclareMathOperator{\Bolza}{{\mathcal B}}
43: 
44: \DeclareMathOperator{\CAT}{{\rm CAT}}
45: 
46: \def\SR{{\rm SR}}
47: 
48: 
49: 
50: \def\go {{\gmetric_{\mathcal O}}} \def\genus {{\it g}}
51: 
52: \def\ie {{\it i.e.\ }} 
53: \def\eg {{\it e.g.\ }} 
54: \def\cf {\hbox{\it cf.\ }}
55: \def\rat{{\angle\!^\mid \! \left(\frac{x}{2}, \frac{\pi}{8}\right)\ }}
56: 
57: \long\def\forget#1\forgotten{} %
58: 
59: \newcommand{\rp}{\mathbb R\mathbb P} 
60: 
61: \numberwithin{equation}{section} 
62: \numberwithin{figure}{section}
63: 
64: \begin{document}
65: 
66: 
67: \author[M.~Katz]{Mikhail G. Katz$^{*}$} \address{Department of
68: Mathematics, Bar Ilan University, Ramat Gan 52900 Israel}
69: \email{katzmik@math.biu.ac.il} \thanks{$^{*}$Supported by the Israel
70: Science Foundation (grants no.\ 620/00-10.0 and 84/03)}
71: 
72: \author[S.~Sabourau]{St\'ephane Sabourau} \address{Laboratoire de
73: Math\'ematiques et Physique Th\'eorique, Universit\'e de Tours, Parc
74: de Grandmont, 37400 Tours, France}
75: 
76: \address{Department of Mathematics, University of Pennsylvania,
77: 209 South 33rd Street, Philadelphia, PA 19104-6395, USA}
78: \email{sabourau@lmpt.math.univ-tours.fr}
79: 
80: \subjclass%[2000] 
81: {Primary 53C20, 53C23 %%Global topological methods (\`a la Gromov) 
82: }
83: 
84: \title[An optimal inequality for CAT(0) metrics in genus two]{An
85: optimal systolic inequality for CAT(0) metrics in genus two}
86: 
87: \keywords{Bolza surface, CAT(0) space, hyperelliptic surface, Voronoi
88: cell, Weierstrass point, systole}
89: 
90: \begin{abstract}
91: We prove an optimal systolic inequality for CAT(0) metrics on a
92: genus~2 surface.  We use a Voronoi cell technique, introduced by
93: C.~Bavard in the hyperbolic context.  The equality is saturated by a
94: flat singular metric in the conformal class defined by the smooth
95: completion of the curve $y^2=x^5-x$.  Thus, among all CAT(0) metrics,
96: the one with the best systolic ratio is composed of six flat regular
97: octagons centered at the Weierstrass points of the Bolza surface.
98: \end{abstract}
99: 
100: \maketitle
101: 
102: \tableofcontents
103: 
104: 
105: 
106: 
107: 
108: 
109: 
110: 
111: 
112: 
113: 
114: 
115: 
116: \section{Hyperelliptic surfaces of nonpositive curvature}
117: \label{Jone1}
118: 
119: 
120: Over half a century ago, a student of C. Loewner's named P. Pu
121: presented, in the pages of the Pacific Journal of Mathematics
122: \cite{Pu}, the first two optimal systolic inequalities, which came to
123: be known as the Loewner inequality for the torus, and Pu's
124: inequality~\eqref{J62} for the real projective plane.
125: 
126: 
127: The recent months have seen the discovery of a number of new systolic
128: inequalities \cite{Am, BK1, Sa1, BK2, IK, BCIK1, BCIK2, KL, Ka4, KS1,
129: KRS}, as well as near-optimal asymptotic bounds \cite{Ka3, KS2, Sa2,
130: KSV, Sa3, RS}.  A number of questions posed in \cite{CK} have thus
131: been answered.  A general framework for systolic geometry in a
132: topological context is proposed in~\cite{KR, KR2}.
133: 
134: The homotopy 1-systole, denoted $\pisys_1(X)$, of a compact metric
135: space $X$ is the least length of a noncontractible loop of $X$.  
136: 
137: 
138: 
139: 
140: 
141: 
142: Given a metric~$\gmetric$ on a surface, let~$\SR(\gmetric)$ denote its
143: systolic ratio
144: \[
145: \SR(\gmetric) =\frac{\pisys_1(\gmetric)^2}{\area(\gmetric)}.
146: \]
147: Given a compact Riemann surface~$\Sigma$, its {\em optimal systolic
148: ratio\/} is defined as~$\SR(\Sigma)= \sup_{\gmetric} \SR(\gmetric)$,
149: where the supremum is over all metrics in the conformal type
150: of~$\Sigma$.  Finally, given a smooth compact surface~$M$, its optimal
151: systolic ratio is defined by setting~$\SR(M)= \sup_\Sigma
152: \SR(\Sigma)$, where the supremum is over all conformal
153: structures~$\Sigma$ on~$M$.  The latter ratio is known for the Klein
154: bottle, \cf \eqref{Jcb}, in addition to the torus and real projective
155: plane already mentioned.
156: 
157: 
158: In the class of all metrics without any curvature restrictions, no
159: singular flat metric on a surface of genus~2 can give the optimal
160: systolic ratio in this genus \cite{Sa1}.  The best available upper
161: bound for the systolic ratio of an arbitrary genus two surface
162: is~$\gamma_2 \simeq 1.1547$, \cf \cite{KS1}.
163: 
164: The precise value of~$\SR$ for the genus~2 surface has so far eluded
165: researchers, \cf~\cite{Ca, Bry}.  We propose an answer in the
166: framework of negatively curved, or more generally,~$\CAT(0)$ metrics.
167: 
168: 
169: The term ``$\CAT(0)$ space'' evokes an extension of the notion of a
170: manifold of nonpositive curvature, to encompass singular spaces.  We
171: will use the term to refer to surfaces with metrics with only mild
172: quotient singularities, defined below.  Here the condition of
173: nonpositive curvature translates into a lower bound of~$2\pi$ for the
174: total angle at the singularity.  We need such an extension so as to
175: encompass the metric which saturates our optimal
176: inequality~\eqref{J11}.
177: 
178: A mild quotient singularity is defined as follows.  Consider a smooth
179: metric on~$\R^2$.  Let~$q\geq 1$ be an integer.  Consider the~$q$-fold
180: cover~$X_q$ of~$\R^2\setminus \{0\}$ with the induced metric.  We
181: compactify~$X_q$ in the neighborhood of the origin to obtain a
182: complete metric space~$X_q^c = X_q \cup \{0\}$.  
183: 
184: \begin{definition}
185: \label{911}
186: Suppose~$X_q^c$ admits an isometric action of~$\Z_p$ fixing the
187: origin.  Then we can form the orbit space~$Y_{p,q} = X_q^c/\Z_p$.  The
188: space~$Y_{p,q}$ is then called mildly singular at the origin.  
189: \end{definition}
190: 
191: The total angle at the singularity is then $\tfrac {2\pi q}{p}$, and
192: the CAT(0) condition is $\tfrac q p \geq 1$.
193: 
194: \begin{remark}
195: \label{tro}
196: Alternatively, a point is singular of total angle $2\pi(1+\beta)$ if
197: the metric is of the form $e^{h(z)} |z|^{2\beta} |dz|^2$ in its
198: neighborhood, where $|dz|^2=dx^2 + dy^2$, \cf
199: M.~Troyanov~\cite[p.~915]{Tro}.
200: \end{remark}
201: 
202: \begin{theorem} 
203: \label{Jtheo:nonpositive}
204: Every~$\CAT(0)$ metric~$\gmetric$ on a surface~$\Sigma_2$ of genus
205: two, satisfies the following optimal inequality:
206: \begin{equation}
207: \label{J11}
208: \SR(\Sigma_2,\gmetric) \leq \tfrac{1}{3} \cot\tfrac{\pi}{8} =
209: \tfrac{1}{3}\left( \sqrt{2} + 1 \right) = 0.8047\ldots
210: \end{equation}
211: The inequality is saturated by a singular flat metric, with~$16$
212: conical singularities, in the conformal class of the Bolza
213: surface.
214: \end{theorem}
215: 
216: The Bolza surface is described in Section~\ref{J22}.  The optimal
217: metric is described in more detail in Section~\ref{Jgenus2}.
218: Theorem~\ref {Jtheo:nonpositive} is proved in Section~\ref{Jproof12}
219: based on the octahedral triangulation of~$S^2$.
220: 
221: \begin{remark}
222: A similar optimal inequality can be proved for hyperelliptic surfaces
223: of genus~5 based on the icosahedral triangulation, \cf~\cite{Bav1}.
224: \end{remark}
225: 
226: 
227: 
228: 
229: 
230: 
231: \section{Distinguishing $16$ points on the Bolza surface}
232: \label{J22}
233: 
234: The Bolza surface~$\Bolza$ is the smooth completion of the affine
235: algebraic curve
236: \begin{equation}
237: \label{J71}
238: y^2= x^5-x.
239: \end{equation}
240: It is the unique Riemann surface of genus two with a group of
241: holomorphic automorphisms of order 48.  (A way of passing from an
242: affine hyperelliptic surface to its smooth completion is described in
243: \cite[p.~60-61]{Mi}.)
244: 
245: \begin{definition}
246: A conformal involution~$J$ of a compact Riemann surface~$\Sigma$ of
247: genus $\genus$ is called {\em hyperelliptic\/} if~$J$ has
248: precisely~$2\genus + 2$ fixed points.  The fixed points of~$J$ are
249: called the Weierstrass points of~$\Sigma$.
250: \end{definition}
251: 
252: The quotient Riemann surface~$\Sigma/J$ is then necessarily the
253: Riemann sphere, denoted henceforth~$S^2$.  Let~$Q: \Sigma \to S^2$ be
254: the conformal ramified double cover, with~$2\genus+2$ branch points.
255: Thus,~$J$ acts on~$\Sigma$ by sheet interchange.  Recall that every
256: surface of genus~2 is hyperelliptic, \ie admits a hyperelliptic
257: involution \cite[Proposition III.7.2]{FK}.
258: 
259: 
260: We make note of $16$ special points on~$\Bolza$.  We refer to a point
261: as being special if it is a fixed point of an order 3 automorphism
262: of~$\Bolza$.
263: 
264: Consider the regular octahedral triangulation of~$S^2=\C\cup \infty$.
265: Its set of vertices is conformal to the set of roots of the polynomial
266: $x^5-x$ of formula~\eqref{J71}, together with the unique point at
267: infinity.  Thus the six points in question can be thought of as the
268: ramification points of the ramified conformal double cover~$Q:{\Bolza}
269: \to S^2$, while the $16$ special points of~$\Bolza$ project to the 8
270: vertices of the cubical subdivision dual to the octahedral
271: triangulation.
272: 
273: In other words, the~$x$-coordinates of the ramification points are
274: \[
275: \left\{ 0, \infty, 1, -1, i, -i \right\},
276: \]
277: which stereographically correspond to the vertices of a regular
278: inscribed octahedron.  The conformal type therefore admits the
279: symmetries of the cube.  If one includes both the hyperelliptic
280: involution, and the real (antiholomorphic) involution of~$\Bolza$
281: corresponding to the complex conjugation~$(x,y)\to (\bar{x}, \bar{y})$
282: of~$\C ^2$, one obtains the full symmetry group~$\Aut(\Bolza)$, of
283: order
284: \begin{equation}
285: \label{J51}
286: |\Aut(\Bolza)| = 96,
287: \end{equation}
288: see \cite[p.~404]{KuN} for more details.
289: 
290: \begin{lemma}
291: \label{J21}
292: The hyperbolic metric of~$\Bolza$ admits~$12$~systolic loops.
293: The~$12$ loops are in one to one correspondence with the edges of the
294: octahedral decomposition of~$S^{2}$.  The correspondence is given by
295: taking the inverse image under~$Q$ of an edge.  The~$12$~systolic
296: loops cut the surface up into~$16$~hyperbolic triangles.  The centers
297: of the triangles are the~$16$~special points.
298: \end{lemma}
299: 
300: See \cite[\S~5]{Sc1} for further details.  The Bolza surface is
301: extremal for two distinct problems:
302: \begin{itemize}
303: \item
304: systole of hyperbolic surfaces \cite{Bav2}, \cite[Theorem 5.2]{Sc1};
305: \item
306: conformal systole of Riemann surfaces \cite{BS}.
307: \end{itemize}
308: The square of the conformal systole of a Riemann surface is also known
309: as its Seshadri constant \cite{Kon}.  The Bolza surface is also
310: conjectured to be extremal for the first eigenvalue of the Laplacian.
311: Such extremality has been verified numerically \cite{JP}.  The above
312: evidence suggests that the systolically extremal surface may lie in
313: the conformal class of~$\Bolza$, as well.  Meanwhile, we have the
314: following result, proved in Section~\ref{Jfive}.
315: 
316: \begin{theorem}
317: \label{J13}
318: The Bolza surface~$\Bolza$ satisfies~$\SR(\Bolza) \leq \frac{\pi}{3}$.
319: \end{theorem}
320: Note that Theorems~\ref{J13} and \ref{Jtheo:nonpositive} imply that
321: $\SR(\Bolza) \in [0.8, 1.05]$.
322: 
323: \section{A flat singular metric in genus~two}
324: \label{Jgenus2}
325: 
326: 
327: The optimal systolic ratio of a genus~2 surface~$(\Sigma_2, \gmetric)$
328: is unknown, but it satisfies the Loewner inequality, \cf \cite{KS1}.
329: Here we discuss a {\em lower\/} bound for the optimal systolic ratio
330: in genus~2, briefly described in~\cite{CK}.
331: 
332: 
333: The example of M. Berger (see \cite [Example~5.6.B$'$]{Gr1}) in
334: genus~2 is a singular flat metric with conical singularities.  It has
335: systolic ratio~$\SR=0.6666$.  This ratio was improved by F. Jenni
336: \cite{Je}, who identified the hyperbolic genus~2 surface with the
337: optimal systolic ratio among all hyperbolic genus~2 surfaces (see also
338: C.~Bavard~\cite{Bav2} and P.~Schmutz \cite[Theorem~5.2]{Sc1}).  The
339: surface in question is a triangle surface~(2,3,8).  Its conformal
340: class is that of the Bolza surface, \cf Section~\ref{J22}.  It admits
341: a regular hyperbolic octagon as a fundamental domain, and has
342: 12~systolic loops of length~$2x$, where~$x=\cosh^{-1} (1+\sqrt{2})$.
343: It has
344: \[
345: \pisys_1= 2\log\left( 1+\sqrt{2}+ \sqrt{2+2\sqrt{2}} \right),
346: \]
347: area~${4\pi}$, and systolic ratio~$\SR\simeq 0.7437$.  This ratio can
348: be improved as follows (see the table of Figure~\ref{Jfig2}).
349: 
350: \begin{figure}
351: \renewcommand{\arraystretch}{1.8} 
352: \[
353: \begin{tabular}[t]{|
354: @{\hspace{3pt}}p{1.5in}|| @{\hspace{3pt}}p{.8in}|
355: @{\hspace{3pt}}p{.8in}| @{\hspace{3pt}}p{.8in}| } \hline & Berger &
356: Jenni & metric $\go$ on Bolza \\ \hline\hline $\displaystyle
357: \SR(\gmetric) =\frac{\pisys_1(\gmetric)^2}{\area(\gmetric)}$ & 0.6666
358: & 0.7437 & 0.8047 \\ \hline
359: \end{tabular}
360: \]
361: \renewcommand{\arraystretch}{1}
362: \caption{\textsf{Three systolic ratios in genus~2}}
363: \label{Jfig2}
364: \end{figure}
365: 
366: 
367: \begin{proposition}
368: The conformal class of the Bolza surface~$\Bolza$ admits a metric,
369: denoted~$\go$, with the following properties:
370: \begin{enumerate}
371: \item
372: the metric is singular flat, with conical singularities precisely at
373: the $16$ special points of Section~$\ref{J22}$;
374: \item
375: each singularity is of total angle~$\frac{9\pi}{4}$, so that the
376: metric~$\go$ is $\CAT(0)$;
377: \item
378: the metric is glued from six flat regular octagons, centered on the
379: Weierstrass points, while the 1-skeleton projects under~$Q:{\Bolza}
380: \to S^2$ to that of the dual cube in~$S^2$;
381: \item
382: the systolic ratio equals~$\SR(\go)= \tfrac{1}{3}\left( \sqrt{2} + 1
383: \right) > \frac{4}{5}$.
384: \end{enumerate}
385: \end{proposition}
386: 
387: 
388: 
389: 
390: 
391: \begin{proof} 
392: The octahedral triangulation of the sphere, \cf Section~\ref{J22},
393: lifts to a triangulation of~$\Bolza$ consisting of $16$ triangles,
394: which we think of as being ``equilateral''.  Here 8 equilateral
395: triangles are connected cyclically around each of the 6 Weierstrass
396: vertices of the triangulation of~$\Bolza$.
397: 
398: We further subdivide each equilateral triangle into 3 isosceles
399: triangles, with a common vertex at the center of the equilateral
400: triangle.  We equip each of the 48 isosceles triangles with a flat
401: metric with obtuse angle $\tfrac{3\pi}{4}$.
402: 
403: 
404: 
405: Each of the 6 Weierstrass vertices of the original triangulation is a
406: smooth point, since the total angle is ~$8\left(\tfrac{\pi}{4} \right)
407: =2\pi$.  Each equilateral triangle possesses a singularity at the
408: center with total angle~$\tfrac{9\pi}{4}=2\pi\left( 1+\tfrac{1}{8}
409: \right)> 2\pi$.  Alternatively, we can apply the Gauss-Bonnet formula
410: $\sum_{\sigma} \alpha (\sigma) = 2\genus-2$, in genus 2 with $16$
411: isometric singularities.  Here the sum is over all
412: singularities~$\sigma$ of a singular flat metric on a surface of
413: genus~$\genus$, where the cone angle at singularity~$\sigma$
414: is~$2\pi(1+\alpha(\sigma))$.  Since the metric~$\go$ is smooth at a
415: Weierstrass point of~$\Bolza$, the metric has
416: only $16$ singularities, precisely at the special points of
417: Section~\ref{J22}, proving items 1 and 2 of the proposition.
418: 
419: Let $x$ denote the sidelength of the equilateral triangle.  The
420: barycentric subdivision of each equilateral triangle consists of 6
421: copies of a flat right angle triangle, denoted $\rat$, with
422: side~$\frac{x}{2}$ and adjacent angle~$\frac{\pi}{8}$.  We thus obtain
423: a decomposition of the metric~$\go$ into 96 copies of the
424: triangle~$\rat$, which can be thought of as a fundamental domain for
425: the action of~$\Aut(\Bolza)$, \cf \eqref{J51}.
426: 
427: 
428: 
429: 
430: 
431: 
432: %The systolic ratio is therefore calculated as follows:
433: %\[
434: %\begin{array}{rcl}
435: %\label{J37}
436: %\displaystyle \frac{\left( \pisys_1(\go)\right)^2} {\area (\go)} & = &
437: %\displaystyle \frac{(2x)^2}{96\;\area\left(\rat \right)}
438: %_{\phantom{a\choose b}} = \displaystyle \frac {x^2}
439: %{24\left(\frac{1}{2} \left( \frac{x} {2} \right)^2 \tan\frac
440: %{\pi}{8}\right)}_{\phantom{a\choose b}} \\ & = & \displaystyle
441: %\left(3\sqrt{3-2\sqrt{2}}\right)^{-1^{\phantom{a}}} \simeq 0.8047,
442: %\end{array}
443: %\]
444: 
445: We have $\pisys_1(\go)= 2x$ by Lemma~\ref{J31}, proving item 4 of the
446: proposition.  To prove item 3, note that the union of the $16$
447: triangles~$\rat$ with a common Weierstrass vertex is a flat regular
448: octagon.  The latter is represented in Figure~\ref{Jfigure1}, together
449: with the systolic loops passing through~it.
450: \end{proof}
451: 
452: \begin{figure}
453: \setlength\unitlength{1pt}
454: \noindent \begin{picture}
455: (400,150)(0,0)
456: \put(120,15){\includegraphics[angle=0,height=120pt]{octagon.eps}}
457: %\put(150,5){{\scshape Figures 9-10}}
458: \end{picture}
459: \caption{\textsf{Flat regular octagon obtained as the union of $16$
460: triangles $\rat$.  The shaded interior octagon represents the region
461: with 4 geodesic loops through every point}}
462: \label{Jfigure1}
463: \end{figure}
464: 
465: 
466: \begin{lemma}
467: \label{J31}
468: The systole of the singular flat $\CAT(0)$ metric on the Bolza
469: surface equals twice the distance between a pair
470: of adjacent Weierstrass points.
471: \end{lemma}
472: 
473: \begin{proof}
474: Consider the smooth closed geodesic $\gamma \subset \Bolza$ which is
475: the inverse image under the map $Q: \Bolza \to S^2$ of an edge of the
476: octahedron, \cf Lemma~\ref{J21}.  Let~$x$ be the distance between a
477: pair of opposite sides of the regular flat octagon in $\Bolza$, or
478: equivalently, the distance between a pair of adjacent Weierstrass
479: points.  Thus, $\length(\gamma)=2x$.  Consider a loop~$\alpha \subset
480: \Bolza$ whose length satisfies
481: \begin{equation}
482: \label{941}
483: \length(\alpha) < 2x.
484: \end{equation}
485: We will prove that there are two possibilities for $\alpha$: it is
486: either contractible, or freely homotopic to one of the 12 geodesics
487: $\gamma$ of the type described above.  On the other hand, $\gamma$ is
488: necessarily length minimizing in its free homotopy class, by the
489: $\CAT(0)$ property of the metric \cite[Theorem~6.8]{BH}.  This will
490: rule out the second possibility, and prove the lemma.
491: 
492: Denote by~$\Bolza^{(1)}$ the graph on~$\Bolza$ given by the inverse
493: image under~$Q$ of the 1-skeleton of the cubical subdivision of~$S^2$.
494: The graph~$\Bolza^{(1)}$ partitions the surface into six regular
495: octagons, denoted $\Omega_k$:
496: \begin{equation}
497: \label{942}
498: \Bolza= \cup_{k=1}^6 \Omega_k.
499: \end{equation}
500: We will deform $\alpha$ to a loop~$\beta \subset \Bolza^{(1)}$, as
501: follows.  The partition \eqref{942} induces a partition of the
502: loop~$\alpha$ into arcs~$\alpha_i$, each lying in its respective
503: octagon $\Omega_{k_i}$.  We deform each~$\alpha_i$, without increasing
504: length, to the straight line segment $[p_i, q_i]\subset \Omega_{k_i}$.
505: The boundary points $p_i, q_i$ of~$\alpha_i$ split the boundary closed
506: curve $\partial \Omega_{k_i} \subset \Bolza^{(1)}$ into a pair of
507: paths.  Let~$\beta_i \subset \partial \Omega_{k_i}$ be the shorter of
508: the two paths.  Denote by~$y$ the distance between adjacent vertices
509: of the octagon.  Then clearly
510: \begin{equation}
511: \label{J32}
512: \length (\beta_i) \leq \frac{4y}{x} \length(\alpha_i).
513: \end{equation}
514: We first deform the loop $\alpha$ into the graph~$\Bolza^{(1)}$.  The
515: deformation fixes the intersection points~$\alpha \cap \Bolza^{(1)}$.
516: Inside $\Omega_{k_i}$, we deform the arc~$\alpha_i$ to the
517: path~$\beta_i$.
518: The length of the resulting loop is at most
519: \[
520: \frac{4y}{x} \length(\alpha) < 8 y
521: \]
522: by \eqref{941} and \eqref{J32}.  Therefore, its homotopy class
523: in~$\Bolza^{(1)}$ can be represented by an imbedded loop $\beta\subset
524: \Bolza^{(1)}$ of length at most~$8y$.  Thus,~$\beta$ contains fewer
525: than 8 edges of $\Bolza^{(1)}$.  Since the number of edges must be
526: even, its image under~$Q$ must retract to a circuit with at most six
527: edges in the 1-skeleton of the cubical subdivision of~$S^2$.  If the
528: number is 4, then the circuit lies in the boundary of a square face of
529: the cube in $S^2$.  But the boundary of a face does not lift
530: to~$\Bolza$, since it surrounds a single ramification point, namely
531: the center of the square face.
532: 
533: Hence there must be 6 edges in the circuit.  There are two types of
534: circuits with six edges in the 1-skeleton of the cubical subdivision
535: of~$S^2$:
536: \begin{enumerate}
537: \item[(a)]
538: the boundary of the union of a pair of adjacent squares; 
539: \item[(b)] a path consisting of the edges meeting a suitable great
540: circle.
541: \end{enumerate}
542: However, a path of type (b) surrounds an odd number, namely 3, of
543: ramification points, and hence does not lift to the genus 2 surface.
544: Meanwhile, a path of type (a) surrounds two ramification points, and
545: hence does lift to the surface.  Such a path is freely homotopic
546: in~$\Bolza$ to one of the 12 geodesics of type
547: $\gamma$ (\cf Lemma~\ref{J21}), completing the proof.
548: \end{proof}
549: 
550: 
551: 
552: 
553: 
554: 
555: 
556: 
557: 
558: \section{Voronoi cells and Euler characteristic}
559: \label{Jproof12}
560: 
561: The following proposition provides a preliminary lower bound on the
562: area of hyperelliptic surfaces with nonpositive curvature.
563: 
564: \begin{proposition} 
565: \label{Jprop:nonpositive}
566: Every~$J$-invariant~$\CAT(0)$ metric~$\gmetric$ on a closed
567: hyperelliptic surface~$\Sigma_\genus$ of genus~$\genus$, satisfies the
568: bound 
569: \[
570: \SR(\Sigma_\genus, \gmetric) \leq {8} ({(\genus+1)\pi})^{-1} .
571: \]
572: \end{proposition}
573: 
574: \begin{proof}
575: To prove this scale-invariant inequality, we normalize the metric
576: on~$\Sigma=\Sigma_\genus$ to unit systole,
577: \ie~$\pisys_1(\Sigma,\gmetric)=1$.  The preimage by~$Q:\Sigma \to S^2$
578: of an arc of~$S^2$ joining two distinct branch points forms a
579: noncontractible loop on~$\Sigma$.  Therefore, the distance between two
580: Weierstrass points is at least~$\frac{1}{2}
581: \pisys_1(\Sigma,\gmetric) = \frac{1}{2}$. Thus, we obtain $2\genus+2$
582: disjoint disks of radius~$\frac{1}{4}$, centered at the Weierstrass
583: points.  Since the metric is~$\CAT(0)$, the area of each disk is at
584: least~$\frac{\pi}{16}$.  Thus,~$\area(\Sigma,\gmetric) \geq
585: \frac{\genus+1}{8} \pi$.
586: \end{proof}
587: 
588: An {\em optimal\/} lower bound requires a more precise estimate on the
589: area of the Voronoi cells.  The idea is to replace area of balls by
590: area of polygons, where control over the number of sides is provided
591: by the Euler characteristic, \cf \cite{Bav2}.
592: 
593: Denote by~$u: \tilde{\Sigma} \rightarrow \Sigma$ its universal cover.  Let
594: $\{x_i \mid i \in \N \}$ be an enumeration of the lifts of Weierstrass
595: points on~$\tilde{\Sigma}$.  The Voronoi cell~$V_i \subset \tilde \Sigma$
596: centered at~$x_i$ is defined as the set of points closer to~$x_i$ than
597: to any other lift of a Weierstrass point.  In formulas,
598: \begin{equation}
599: \label{J}
600: V_i = \left\{ \left. x \in \tilde{\Sigma} \right| d(x,x_i) \leq d(x,x_j)
601: \text{ for every } j \neq i \right\}.
602: \end{equation}
603: The Voronoi cells on~$\tilde{\Sigma}$ are polygons whose edges are arcs of
604: the equidistant curves between a pair of lifts of Weierstrass
605: points. Note that these edges are not
606: necessarily geodesics.  The Voronoi cells on~$\tilde{\Sigma}$ are
607: topological disks, while their projections~$u(V_i)\subset \Sigma$ may have
608: more complicated topology.  Thus, the surface~$\Sigma$ decomposes
609: into~$2\genus+2$ images of Voronoi cells, centered at the~$2\genus+2$
610: Weierstrass points.  By the number of sides of~$u(V)$ we will mean the
611: number of sides of the polygon~${V}$.
612: 
613: 
614: 
615: \begin{lemma}
616: Let $\gmetric_{1}$ and $\gmetric_{2}$ be two $\CAT(0)$ metrics lying
617: in the same conformal class.  Then, the averaged metric
618: $\gmetric=\tfrac{1}{2} (\gmetric_{1}+\gmetric_{2})$ is $\CAT(0)$, as
619: well.
620: \end{lemma}
621: 
622: \begin{proof}
623: Choose a point~$x \in \Sigma$ and a metric~$\gmetric_{0}$ in its
624: conformal class, which is flat in a neighborhood of~$x$.  Every metric
625: $\gmetric = H \gmetric_{0}$ conformal to it satisfies
626: \[
627: K_{\gmetric} H = K_{\gmetric_{0}} - \tfrac{1}{2} \Delta \log H,
628: \]
629: \cf \cite[p.~265]{GHL}, where $K_{\gmetric}$ and $K_{\gmetric_{0}}$
630: are the Gaussian curvatures of $\gmetric$ and~$\gmetric_{0}$, and
631: $\Delta$ is the Laplacian of~$\gmetric_{0}$ with $\Delta f =
632: \mbox{div} \nabla f$.  Thus, the metrics $\gmetric_{i}$ can be written
633: as $\gmetric_{i} = e^{h_{i}} \gmetric_{0}$ where $h_{i}$ is
634: subharmonic in the neighborhood of~$x$, \ie $\Delta h_{i} \geq 0$.  A
635: simple computation shows that
636: \[
637: \Delta \log H \geq \frac{e^{h_{1}} \Delta h_{1} + e^{h_{2}} \Delta
638: h_{2}}{2H} \geq 0
639: \]
640: where $H = \tfrac{1}{2} (e^{h_{1}} + e^{h_{2}})$, proving the lemma if
641: both points are regular.  For singular points with positive angle
642: excess, the CAT(0) property for the averaged metric is immediate from
643: Remark~\ref{tro}.
644: \end{proof}
645: 
646: \begin{proof}[Proof of Theorem~\ref{Jtheo:nonpositive}]
647: Since averaging by $J$ can only improve the systolic ratio, we may
648: assume without loss of generality that our metric is already
649: $J$-invariant.
650: 
651: \forget
652: Consider the exponential map~$\exp_A$ at a Weierstrass point~$A\in \Sigma$.
653: If~$A$ is a singular point, we lift to the appropriate smooth finite
654: cover, and use its exponential map, \cf Definition~\ref{911}.
655: 
656: Let~$B\in \Sigma$ be another Weierstrass point.  Consider its lift~$B_0 =
657: \exp_A^{-1}(B)$ to the tangent plane~$T_A$.  Consider the equidistant
658: line
659: \[ 
660: L_{O, B_0} \subset T_A
661: \]
662: between the origin~$O\in T_A$ and the point~$B_0$.
663: 
664: By the Toponogov theorem applied to the hinge at~$O$ defined by the
665: point~$B_0$ and an arbitrary point~$X\in L_{O, B_0}$ on the
666: equidistant line, we have
667: \begin{equation}
668: \label{J44b}
669: \dist_\Sigma(A, \exp_A(X)) \leq \dist_\Sigma(B, \exp_A(X)) .
670: \end{equation}
671: \forgotten
672: 
673: 
674: 
675: There exists an extension of the notions of tangent plane and
676: exponential map, to surfaces with singularities.  Namely, let $A\in
677: \Sigma$.  There exists a CAT(0) piecewise flat plane~$T_{A}$ with
678: conical singularities and a covering $\exp_{A}:T_{A} \to \Sigma$ with
679: the following properties:
680: \begin{enumerate}
681: \item 
682: $\exp_A$ sends the origin $O$ of $T_{A}$ to~$A$;
683: 
684: \item 
685: $\exp_A$ takes the conical singularities of~$T_{A}$ to the
686: singularities of~$\Sigma$;
687: 
688: \item 
689: $\exp_A$ sends every pair of geodesic arcs issuing from the origin $O
690: \in T_{A}$ and forming an angle~$\theta$, to a pair of geodesic arcs
691: of the same lengths, and forming the same angle, at their
692: basepoint~$A$.
693: \end{enumerate}
694: 
695: By the Rauch Comparison Theorem, the exponential map $\exp_{A}$ does
696: not decrease distances.
697: 
698: Now assume $A\in \Sigma$ is a Weierstrass point, and let~$B\in \Sigma$
699: be another Weierstrass point.  Fix a lift~$B_0$ of~$B$ to the tangent
700: plane~$T_A$, along a minimizing arc.  Consider the equidistant line
701: \[ 
702: L_{O, B_0} \subset T_A
703: \]
704: between the origin~$O\in T_A$ and the point~$B_0$.  Consider a
705: point~$X_{0}\in L_{O, B_0}$.  Let $X=\exp_{A}(X_{0})$.  Since the
706: exponential map does not decrease distances, we have
707: \begin{equation}
708: \label{J44b}
709: \dist_\Sigma(A, X) = \dist_{T_{A}}(O,X_{0}) = \dist_{T_{A}} (B_{0},
710: X_{0}) \leq \dist_\Sigma(B, X) .
711: \end{equation}
712: 
713: Now consider the polygon in the tangent plane~$T_{A}$, obtained as the
714: intersection of the halfspaces containing the origin, defined by the
715: lines~$L_{O, B_0}$, as~$B$ runs over all Weierstrass points.  It
716: follows from~\eqref{J44b} that the exponential image of this polygon
717: is contained in the Voronoi cell of~$A$.  Since the exponential map
718: does not decrease distances, the area of the polygon is a lower bound
719: for the area of the Voronoi cell.  If~$k$ is the number of sides
720: of~$V$, then~$V$ is partitioned into~$k$ triangles with
721: angle~$\theta_i$ at~$O$, whose area is bounded below by
722: \begin{equation}
723: \label{J43}
724: \left( \tfrac{\pisys_1}{4} \right)^2 \tan\tfrac{\theta_i}{2}
725: \end{equation}
726: since $\dist(A,B) \geq \frac{\pisys_{1}}{2}$.
727: 
728: Consider the graph on~$S^2$ defined by the projections of the Voronoi
729: cells to the sphere.  Thus we have~$f=6$ faces.  Applying the formula
730: $v-e+f=2$ and the well-known fact that~$3v \leq {2} e$, we obtain
731: \[
732: {e} \leq 3f-6 =12.
733: \]
734: Hence the spherical graph has at most~12 edges.  Note that the maximum
735: is attained by the 1-skeleton of the cubical subdivision.
736: 
737: The area of a flat isosceles triangle with third angle~$\theta$, and
738: with unit altitude from the third vertex, is~$\tan \tfrac {\theta}
739: {2}$.  This formula provides a lower bound for the area of the Voronoi
740: cells as in \eqref{J43}.  The proof is completed by Jensen's
741: inequality applied to the convex function~$\tan (\tfrac {x} {2})$
742: when~$0 < x < \pi$.  In the boundary case of equality, we have~$e=12$,
743: all angles~$\theta_i$ as in~\eqref{J43} must be equal, curvature must
744: be zero because of equality in the Rauch Comparison Theorem, and we easily deduce that
745: each Voronoi cell is a regular octagon.  To minimize the area of the
746: octagon, we must choose~$\theta$ as small as possible.  The~$\CAT(0)$
747: hypothesis at the center of the octagon imposes a lower bound~$\theta
748: \geq \tfrac \pi{4}$.  Hence the optimal systolic ratio is achieved for
749: the regular flat octagon with a smooth point at the center.
750: \end{proof}
751: 
752: 
753: \section{Arbitrary metrics on the Bolza surface}
754: \label{Jfive}
755: 
756: The conformal class of the Bolza surface~$\Bolza$, \cf
757: Section~\ref{J22}, is likely to contain a systolically optimal surface
758: in genus~2, as discussed in Section~\ref{Jone1}.
759: 
760: 
761: 
762: 
763: 
764: \begin{theorem} \label{theo:conf}
765: Every metric~$\gmetric$ in the conformal class of the Bolza surface
766: satisfies the bound
767: \begin{equation}
768: \label{J12}
769: \SR(\gmetric) \leq \tfrac{\pi}{3} = 1.0471\ldots
770: \end{equation}
771: \end{theorem}
772: 
773: \begin{remark}
774: In particular, every metric~$\gmetric$ in the conformal class of the
775: Bolza surface satisfies Bavard's inequality
776: \begin{equation}
777: \label{Jcb}
778: \SR(\gmetric) \leq \frac{\pi}{2^{3/2}} \simeq 1.1107
779: \end{equation}
780: for the Klein bottle \cite{Bav1}.  This suggests a possible
781: monotonicity of~$\chi(\Sigma)$ as a function of~$\SR(\Sigma)$.
782: \end{remark}
783: 
784: \begin{lemma}
785: \label{J54}
786: Let~$\gmetric$ be an~$\Aut(\Bolza)$-invariant metric on~$\Bolza$.
787: Consider the displacement~$\delta(J)$ of~$J$ on the
788: 1-skeleton~$\Bolza^{(1)}$ of the Voronoi subdivision of~$\Bolza$,
789: namely
790: \[
791: \delta(J) = \min_{x\in \Bolza^{(1)}} \dist(x, J(x)).
792: \]
793: Then~$\area(\Bolza,\gmetric) \geq 6 \left( \tfrac{2}{\pi} \right)
794: \delta(J)^2$.
795: \end{lemma}
796: 
797: 
798: 
799: \begin{proof}
800: Consider the Voronoi subdivision with respect to the set of six
801: Weierstrass points on~$\Bolza$.  Since each Voronoi
802: cell~$\Omega\subset \Bolza$ is~$J$-invariant, we can identify all
803: pairs of opposite points of the boundary~$\partial \Omega$, to obtain
804: a projective plane
805: \begin{equation}
806: \label{J72} 
807: \rp^2 = \Omega/\sim.
808: \end{equation}
809: We now apply Pu's inequality \cite{Pu} to each of the six Voronoi
810: cells, to obtain
811: \begin{equation}
812: \label{J62}
813: \area(\rp ^2)\geq \tfrac{2}{\pi} \pisys_1(\rp^2)^2 .
814: \end{equation}
815: The lemma now follows from the bound~$\pisys_1(\rp^2) \geq \delta(J)$.
816: \end{proof}
817: 
818: \begin{lemma}
819: \label{J74}
820: Every~$\Aut(\Bolza)$-invariant metric~$\gmetric$ on~$\Bolza$ satisfies
821: the bound $2 \delta(J) \geq \pisys_1(\Bolza, \gmetric)$.
822: \end{lemma}
823: 
824: \begin{proof}
825: Let~$p, J(p)\in \partial \Omega$ satisfy~$\dist(p, J(p)) = \delta(J)$.
826: Let~$\alpha \subset \Omega$ be a minimizing path joining~$p$
827: to~$J(p)$.  Let~$\Omega'\subset \Bolza$ be the adjacent Voronoi cell
828: containing this pair of boundary points, and~$r: \Bolza \to \Bolza$
829: the anticonformal involution which switches~$\Omega$ and~$\Omega'$,
830: and fixes their common boundary.  The loop~$\alpha \cup r(\alpha)$
831: belongs to the free homotopy class of the noncontractible loop~$\gamma
832: \subset \Bolza$ obtained as the inverse image under~$Q: \Bolza \to
833: S^2$ of the edge of the octahedral decomposition of~$S^{2}$ joining
834: the images of the centers of~$\Omega$ and~$\Omega'$, \cf
835: Lemma~\ref{J21}.  Since~$\length(\alpha \cup r(\alpha)) = 2
836: \delta(J)$, the lemma follows.
837: \end{proof}
838: 
839: 
840: \begin{proof}[Proof of Theorem~\ref{theo:conf}]
841: We may assume that the metric on~$\Bolza$ is~$\Aut(\Bolza)$-invariant,
842: since averaging the metric by a finite group of holomorphic and
843: antiholomorphic diffeomorphisms can only improve the systolic ratio.
844: We combine the inequalities of Lemmas~\ref{J74} and \ref{J54} to prove
845: the theorem.
846: \end{proof}
847: 
848: 
849: 
850: %\vfill\eject
851: \begin{thebibliography}{ABCDe}
852: 
853: \bibitem[Am04]{Am} Ammann, B.: Dirac eigenvalue estimates on two-tori.
854: {\em J. Geom. Phys.} \textbf{51} (2004), no. 3, 372--386.
855: 
856: 
857: \bibitem[BCIK05]{BCIK1} Bangert, V; Croke, C.; Ivanov, S.; Katz, M.:
858: Filling area conjecture and ovalless real hyperelliptic surfaces, {\em
859: Geometric and Functional Analysis (GAFA)\/} \textbf{25} (2005), no.~3.
860: 577-597.  See \texttt{arXiv:math.DG/0405583}
861: 
862: \bibitem[BCIK06]{BCIK2} Bangert, V; Croke, C.; Ivanov, S.; Katz, M.:
863: Boundary case of equality in optimal Loewner-type inequalities, {\em
864: Trans. Amer. Math. Soc.} \textbf{358} (2006).  See
865: \texttt{arXiv:math.DG/0406008}
866: 
867: 
868: 
869: \bibitem[BK03]{BK1} Bangert, V.; Katz, M.: Stable systolic
870: inequalities and cohomology products, {\em Comm. Pure Appl. Math.}
871: \textbf{56} (2003), 979--997.  Available at
872: \texttt{arXiv:math.DG/0204181}
873: 
874: \bibitem[BK04]{BK2} Bangert, V; Katz, M.: An optimal Loewner-type
875: systolic inequality and harmonic one-forms of constant norm.  {\em
876: Comm. Anal. Geom.}  \textbf{12} (2004), number 3, 703-732.  See
877: \texttt{arXiv:math.DG/0304494}
878: 
879: \bibitem[Bav86]{Bav1} Bavard, C.: In\'egalit\'e isosystolique pour la
880: bouteille de Klein.  {\em Math. Ann.} \textbf{274} (1986), no. 3,
881: 439--441.
882: 
883: 
884: \bibitem[Bav92]{Bav2} Bavard, C.: La systole des surfaces
885: hyperelliptiques, {\em Prepubl. Ec. Norm. Sup. Lyon\/} \textbf{71}
886: (1992).
887: 
888: 
889: 
890: \bibitem[BH99]{BH} Bridson, M.; Haefliger, A.: Metric spaces of
891: non-positive curvature.  {\em Grundlehren der Mathematischen
892: Wissenschaften}, \textbf{319}.  Springer-Verlag, Berlin, 1999.
893: 
894: 
895: 
896: 
897: \bibitem[Br96]{Bry} Bryant, R.: On extremals with prescribed
898: Lagrangian densities, Manifolds and geometry (Pisa, 1993), 86--111,
899: {\em Sympos. Math.}, \textbf{36}, Cambridge Univ. Press, Cambridge,
900: 1996.
901: 
902: \bibitem[BS94]{BS} Buser, P.; Sarnak, P.: On the period matrix of a
903: Riemann surface of large genus.  With an appendix by J. H. Conway and
904: N. J. A. Sloane.  {\em Invent. Math.} \textbf{117} (1994), no.~1,
905: 27--56.
906: 
907: \bibitem[Ca96]{Ca} Calabi, E.: Extremal isosystolic metrics for
908: compact surfaces, Actes de la table ronde de geometrie differentielle,
909: {\em Sem. Congr.}  \textbf{1} (1996), Soc. Math. France, 167--204.
910: 
911: 
912: \bibitem[CK03]{CK} Croke, C.; Katz, M.: Universal volume bounds in
913: Riemannian mani\-folds, {\it Surveys in Differential Geometry\/}
914: \textbf{8}, Lectures on Geometry and Topology held in honor of Calabi,
915: Lawson, Siu, and Uhlenbeck at Harvard University, May 3- 5, 2002,
916: edited by S.T. Yau (Somerville, MA: International Press, 2003.)
917: pp. 109 - 137.  See \texttt{arXiv:math.DG/0302248}
918: 
919: 
920: \bibitem[FK92]{FK} Farkas, H.; Kra, I.: Riemann surfaces.  Second
921: edition. Graduate Texts in Mathematics, 71. Springer-Verlag, New York,
922: 1992.
923: 
924: \bibitem[GHL90]{GHL} Gallot, S.; Hulin, D.; Lafontaine, J.: Riemannian
925: Geometry. Springer-Verlag, Berlin, 1990.
926: 
927: \bibitem[Gr83]{Gr1} Gromov, M.: Filling Riemannian manifolds, {\em
928: J. Diff. Geom.\/} \textbf{18} (1983), 1-147.
929: 
930: \bibitem[Gr96]{Gr2} Gromov, M.: Systoles and intersystolic
931: inequalities.  Actes de la Table Ronde de G\'{e}om\'{e}trie
932: Diff\'{e}rentielle (Luminy, 1992), 291--362, {\em S\'{e}min. Congr.}
933: \textbf{1}, Soc. Math. France, Paris, 1996.
934: \newline\noindent
935: \texttt{www.emis.de/journals/SC/1996/1/ps/smf\_sem-cong\_1\_291-362.ps.gz}
936: 
937: \bibitem[Gr99]{Gr3} Gromov, M.: Metric structures for Riemannian and
938: non-Riemannian spaces.  {\em Progr. in Mathematics}, \textbf{152},
939: Birkh\"{a}user, Boston, 1999.
940: 
941: \bibitem[IK04]{IK} Ivanov, S.; Katz, M.: Generalized degree and
942: optimal Loewner-type inequalities, {\em Israel J. Math.}  \textbf{141}
943: (2004), 221-233.  \texttt{arXiv:math.DG/0405019}
944: 
945: 
946: \bibitem[JP04]{JP} Jakobson, D.; Levitin, M.; Nadirashvili, N.;
947: Polterovich, I.: Spectral problems with mixed Dirichlet-Neumann
948: boundary conditions: isospectrality and beyond, preprint.  See
949: \texttt{arXiv:math.SP/0409154}
950: 
951: \bibitem[Je84]{Je} Jenni, F.: \"Uber den ersten Eigenwert des
952: Laplace-Operators auf ausgew\"{a}hlten Beispielen kompakter
953: Riemannscher Fl\"{a}chen. [On the first eigenvalue of the Laplace
954: operator on selected examples of compact Riemann surfaces] {\em
955: Comment. Math. Helv.} \textbf{59} (1984), no. 2, 193--203.
956: 
957: 
958: \bibitem[Ka03]{Ka3} Katz, M.: Four-manifold systoles and surjectivity
959: of period map, {\em Comment. Math. Helv.} \textbf{78} (2003), 772-786.
960: \texttt{arXiv:math.DG/0302306}
961: 
962: \bibitem[Ka06]{Ka4} Katz, M.: Systolic inequalities and Massey
963: products in simply-connected manifolds.  See
964: \texttt{arXiv:math.DG/0604012}
965: 
966: 
967: \bibitem[KL05]{KL} Katz, M.; Lescop, C.: Filling area conjecture,
968: optimal systolic inequalities, and the fiber class in abelian covers.
969: Geometry, spectral theory, groups, and dynamics, 181--200, {\em
970: Contemp. Math.} \textbf{387}, Amer. Math. Soc., Providence, RI, 2005.
971: See \texttt{arXiv:math.DG/0412011}
972: 
973: 
974: \bibitem[KR06]{KR} Katz, M.; Rudyak, Y.: Lusternik-Schnirelmann
975: category and systolic category of low dimensional manifolds.  {\em
976: Communications on Pure and Applied Mathematics}, \textbf{59} (2006).
977: See \texttt{arXiv:math.DG/0410456}
978: 
979: \bibitem[KR07]{KR2} Katz, M.; Rudyak, Y.: Bounding volume by systoles
980: of 3-manifolds.  See \texttt{arXiv:math.DG/0504008}
981: 
982: 
983: \bibitem[KRS07]{KRS} Katz, M.; Rudyak, Y.: Sabourau, S.: Systoles of
984: 2-complexes, Reeb graph, and Grushko decomposition.  See
985: \texttt{arXiv:math.DG/0602009}
986: 
987: 
988: \bibitem[KS05]{KS2} Katz, M.; Sabourau, S.: Entropy of systolically
989: extremal surfaces and asymptotic bounds, {\em Ergodic Theory and
990: Dynamical Systems\/} \textbf{25} (2005), no.~4, 1209-1220.  Available
991: at \texttt{arXiv:math.DG/0410312}
992: 
993: \bibitem[KS06]{KS1} Katz, M.; Sabourau, S.: Hyperelliptic surfaces are
994: Loewner, {\em Proc. Amer. Math. Soc.} \textbf{134} (2006), no.~4,
995: 1189-1195.  See \texttt{arXiv:math.DG/0407009}
996: 
997: \bibitem[KSV06]{KSV} Katz, M.; Schaps, M.; Vishne, U.: Logarithmic
998: growth of systole of arithmetic Riemann surfaces along congruence
999: subgroups.  Available at \texttt{arXiv:math.DG/0505007}
1000: 
1001: \bibitem[Kon03]{Kon} Kong, J.: Seshadri constants on Jacobian of
1002: curves.  {\em Trans. Amer. Math. Soc.} \textbf{355} (2003), no. 8,
1003: 3175--3180.
1004: 
1005: \bibitem[KuN95]{KuN} Kuusalo, T.; N\"a\"at\"anen, M.: Geometric
1006: uniformization in genus~$2$.  {\em Ann. Acad. Sci. Fenn. Ser. A Math.}
1007: \textbf{20} (1995), no. 2, 401--418.
1008: 
1009: \bibitem[Mi95]{Mi} Miranda, R.: Algebraic curves and Riemann
1010: surfaces. {\em Graduate Studies in Mathematics}, {\textbf 5}.
1011: Amer. Math. Soc., Providence, RI, 1995.
1012: 
1013: \bibitem[Pu52]{Pu} Pu, P.M.: Some inequalities in certain
1014: nonorientable Riemannian manifolds, {\it Pacific J. Math.\/}
1015: \textbf{2} (1952), 55--71.
1016: 
1017: 
1018: \bibitem[RS07]{RS} Rudyak, Y.: Sabourau, S.: Systolic invariants of
1019: groups and $2$-complexes via Grushko decomposition.
1020: 
1021: \bibitem[Sa04]{Sa1} Sabourau, S.: Systoles des surfaces plates
1022: singuli\`eres de genre deux, {\em Math. Zeitschrift\/} \textbf{247}
1023: (2004), no. 4, 693--709.
1024: 
1025: \bibitem[Sab06]{Sa2} Sabourau, S.: Systolic volume and minimal entropy
1026: of aspherical manifolds, {\em J. Differential Geom.}, to appear.  See
1027: \texttt{arXiv:math.DG/0603695}
1028: 
1029: \bibitem[Sa06b]{Sa3} Sabourau, S.: Asymptotic bounds for separating
1030: systoles on surfaces, preprint.
1031: 
1032: \bibitem[Sc93]{Sc1} Schmutz, P.: Riemann surfaces with shortest
1033: geodesic of maximal length.  {\em Geom. Funct. Anal.} \textbf{3}
1034: (1993), no. 6, 564--631.
1035: 
1036: \bibitem[Tr90]{Tro} Troyanov, M.: Coordonn\'ees polaires sur les
1037: surfaces riemanniennes singuli\`eres.  {\em Ann. Inst. Fourier
1038: (Grenoble)} \textbf{40} (1990), no. 4, 913--937 (1991).
1039: 
1040: \end{thebibliography}
1041: 
1042: 
1043: 
1044: 
1045: \end{document}
1046: 
1047: 
1048: 
1049: