math0204007/fat.tex
1: \documentclass[twocolumn,amsmath,amssymb,aps,secnumarabic,nofootinbib,%
2:     superscriptaddress,floatfix]{revtex4}
3: \usepackage{amsthm,times,mathptmx,graphicx,pstricks,pst-node}
4: \bibliographystyle{hamsplain}
5: 
6: \newtheorem{theorem}{Theorem}
7: \newtheorem{lemma}[theorem]{Lemma}
8: \newtheorem{corollary}[theorem]{Corollary}
9: \newtheorem{proposition}[theorem]{Proposition}
10: \newtheorem{question}[theorem]{Question}
11: \theoremstyle{remark}
12: \newtheorem*{remark}{Remark}
13: \newcommand{\Hom}{\operatorname{Hom}}
14: \newcommand{\conv}{\operatorname{conv}}
15: \newcommand{\Z}{\mathbb{Z}}
16: \newcommand{\R}{\mathbb{R}}
17: \newcommand{\F}{\mathbb{F}}
18: \newcommand{\HH}{\mathbb{H}}
19: \renewcommand{\hat}{\widehat}
20: \renewcommand{\tilde}{\widetilde}
21: \renewcommand{\d}{\partial}
22: \newcommand{\st}{\mathrm{st}}
23: \newcommand{\hg}{{\hat{g}}}
24: \newcommand{\hS}{{\hat{S}}}
25: \newcommand{\pol}{{}^\triangle}
26: \newcommand{\ppol}{{}^\diamondsuit}
27: \newcommand{\maxfatness}{5.048}
28: \newcommand{\eqdef}{\stackrel{\mathrm{def}}{=}}
29: 
30: \newcommand{\eg}{\emph{e.g.}}
31: \newcommand{\Eg}{\emph{E.g.}}
32: \newcommand{\ie}{\emph{i.e.}}
33: \newcommand{\cf}{\emph{cf.}}
34: \newcommand{\eatline}{\vspace{-\baselineskip}}
35: \newcommand{\blankline}{\vspace{\baselineskip}}
36: 
37: \renewcommand{\figurename}{Figure}
38: \newenvironment{fullfigure}[2]
39:     {\begin{figure}[ht]\begin{center}\def\ffa{#1}\def\ffb{#2}}
40:     {\vspace{\baselineskip}\caption{\ffb.}\label{\ffa}\end{center}\end{figure}}
41: \renewcommand{\tablename}{Table}
42: \renewcommand{\thetable}{\arabic{table}}
43: 
44: \newcommand{\fig}[1]{Figure~\ref{#1}}
45: \newcommand{\thm}[1]{Theorem~\ref{#1}}
46: \renewcommand{\sec}[1]{Section~\ref{#1}}
47: \newcommand{\lem}[1]{Lemma~\ref{#1}}
48: \newcommand{\prop}[1]{Proposition~\ref{#1}}
49: \newcommand{\eq}[2]{\begin{equation}\label{#1}#2\end{equation}}
50: 
51: \psset{linewidth=.5pt,dash=2.5pt 2.5pt,unit=.25in,arrowsize=4pt 3}
52: \SpecialCoor
53: 
54: \newgray{gray90}{.90}
55: \newcommand{\ar}{\pspicture(0,0)(0,0)
56:     \psline[arrows=->](4.3pt,0)(4.4pt,0)\endpspicture}
57: \newcommand{\sgpie}{
58:     \pscustom[linestyle=none,fillstyle=solid,fillcolor=gray90]{
59:     \psarcn(4.899;30){1.464}{255}{165} \psarcn(4.899;60){1.464}{285}{195}
60:     \psarcn(4.899;90){1.464}{315}{225} \psarcn(4.899;120){1.464}{345}{255}
61:     \psarcn(4.899;150){1.464}{15}{285} \psarcn(4.899;210){1.464}{75}{345}
62:     \psarcn(4.899;240){1.464}{105}{15} \psarcn(4.899;270){1.464}{135}{45}
63:     \psarcn(4.899;300){1.464}{165}{75} \psarcn(4.899;330){1.464}{195}{105}}
64:     \psarcn(4.899;30){1.464}{255}{165} \psarcn(4.899;60){1.464}{285}{195}
65:     \psarcn(4.899;90){1.464}{315}{225} \psarcn(4.899;120){1.464}{345}{255}
66:     \psarcn(4.899;150){1.464}{15}{285} \psarcn(4.899;210){1.464}{75}{345}
67:     \psarcn(4.899;240){1.464}{105}{15} \psarcn(4.899;270){1.464}{135}{45}
68:     \psarcn(4.899;300){1.464}{165}{75} \psarcn(4.899;330){1.464}{195}{105}
69:     \qdisk(4;165){.1} \qdisk(4;195){.1} \qdisk(4;225){.1} \qdisk(4;255){.1}
70:     \qdisk(4;285){.1} \qdisk(4;315){.1} \qdisk(4;345){.1} \qdisk(4;15){.1}
71:     \qdisk(4;45){.1} \qdisk(4;75){.1} \qdisk(4;105){.1} \qdisk(4;135){.1}
72:     \rput{120}(3.435;30){\ar} \rput{150}(3.435;60){\ar}
73:     \rput{180}(3.435;90){\ar} \rput{210}(3.435;120){\ar}
74:     \rput{240}(3.435;150){\ar} \rput{300}(3.435;210){\ar}
75:     \rput{330}(3.435;240){\ar} \rput{0}(3.435;270){\ar}
76:     \rput{30}(3.435;300){\ar} \rput{60}(3.435;330){\ar}
77:     \uput[180](4;180){\large $\vdots$}
78:     \uput[0](4;0){\large $\vdots$}}
79: 
80: \begin{document}
81: \title{Fat 4-polytopes and fatter 3-spheres}
82: 
83: \author{David Eppstein}
84: \email[Email: ]{eppstein@ics.uci.edu}
85: \thanks{Supported in part by NSF grant CCR \#9912338}
86: \affiliation{Department of Information and Computer Science,
87:     University of California, Irvine, CA 92697}
88: \author{Greg Kuperberg}
89: \email[Email: ]{greg@math.ucdavis.edu}
90: \thanks{Supported by NSF grant DMS \#0072342}
91: \affiliation{Department of Mathematics,
92:     University of California, Davis, CA 95616}
93: \author{G\"unter M. Ziegler}
94: \email[Email: ]{ziegler@math.tu-berlin.de}
95: \thanks{Supported by a DFG Leibniz grant}
96: \affiliation{Institut f\"ur Mathematik, MA 6-2,
97:     Technische Universit\"at Berlin, D-10623 Berlin, Germany}
98: 
99: \begin{abstract}
100: We introduce the \emph{fatness} parameter of a $4$-dimensional polytope $P$,
101: defined as $\phi(P)=(f_1+f_2)/(f_0+f_3)$.  It arises in an important open
102: problem in $4$-dimensional combinatorial geometry: Is the fatness of convex
103: $4$-polytopes bounded?
104: 
105: We describe and analyze a hyperbolic geometry construction that produces
106: $4$-polytopes with fatness $\phi(P)>\maxfatness$, as well as the first infinite
107: family of $2$-simple, $2$-simplicial $4$-polytopes. Moreover, using a
108: construction via finite covering spaces of surfaces, we show that fatness is
109: not bounded for the more general class of strongly regular CW decompositions of
110: the $3$-sphere.
111: \end{abstract}
112: 
113: \maketitle
114: 
115: \section{Introduction}
116: \label{s:intro}
117: 
118: The characterization of the set ${\cal F}_3$ of $f$-vectors of convex 3-dimensional
119: polytopes (from 1906, due to Steinitz \cite{Steinitz:eulerschen}) is well-known
120: and explicit, with a simple proof: An integer vector $(f_0,f_1,f_2)$ is the
121: $f$-vector of a $3$-polytope if and only if it satisfies
122: \begin{itemize}\setlength{\itemsep}{0pt}
123: \item $f_1=f_0+f_2-2$ (the Euler equation),
124: \item $f_2\le 2f_0-4$ (with equality for simplicial polytopes), and
125: \item $f_0\le 2f_2-4$ (with equality for simple polytopes).
126: \end{itemize}
127: (Recall that by the definition of the $f$-vector, $f_k$ is the number
128: of $k$-faces of the polytope.)
129: This simple result is interesting for several reasons:
130: \begin{itemize}\setlength{\itemsep}{0pt}
131: \item The set of $f$-vectors is the set of \emph{all} the integer points in a
132: closed $2$-dimensional \emph{polyhedral cone} (whose apex is the $f$-vector
133: $f(\Delta_3)=(4,6,4)$ of a $3$-dimensional simplex). In particular, it is
134: \emph{convex} in the sense that ${\cal F}_3=\conv({\cal F}_3)\cap \Z^3$.
135: \item The same characterization holds for convex $3$-polytopes (geometric
136: objects), more generally for strongly regular CW $2$-spheres (topological
137: objects), and yet more generally for Eulerian lattices of length $4$
138: (combinatorial objects \cite{Stanley:eulerian}).
139: \end{itemize}
140: 
141: In contrast to this explicit and complete description of ${\cal F}_3$, our
142: knowledge of the set ${\cal F}_4$ of $f$-vectors of (convex) $4$-polytopes (see
143: Bayer \cite{Bayer:extended} and H\"oppner and Ziegler \cite{HZ:census}) is very
144: incomplete. We know that the set ${\cal F}_4$ of all $f$-vectors of
145: 4-dimensional polytopes has no similarly simple description. In particular, the
146: convex hull of ${\cal F}_4$ is not a cone, it is not a closed set, and not all
147: integer points in the convex hull are $f$-vectors. Also, the $3$-dimensional
148: cone with apex $f(\Delta_4)$ spanned by ${\cal F}_4$ is not closed, and its
149: closure may not be polyhedral.
150: 
151: Only the two extreme cases of simplicial and of simple $4$-polytopes (or
152: $3$-spheres) are well-understood.   Their $f$-vectors correspond to faces of
153: the convex hull of ${\cal F}_4$, defined by the valid inequalities $f_2\ge
154: 2f_3$ and $f_1\ge 2f_0$, and the $g$-Theorem, proved for $4$-polytopes by
155: Barnette \cite{Barnette:inequalities} and for $3$-spheres by Walkup
156: \cite{Walkup:lower}, provides \emph{complete} characterizations of their
157: $f$-vectors. (The $g$-Theorem for general simplicial polytopes was famously
158: conjectured by McMullen \cite{McMullen:simplicial} and proved by Billera and
159: Lee \cite{BL:sufficiency} and Stanley \cite{Stanley:simplicial}.  See
160: \cite[\S8.6]{Ziegler:gtm} for a review.)
161: 
162: But we have no similarly complete picture of other extremal types of
163: $4$-polytopes. In particular, we cannot currently answer the following key
164: question: Is there a constant $c$ such that all  $4$-dimensional convex
165: polytopes $P$ satisfy the inequality
166: \[
167: f_1(P)+f_2(P)\;\le\;c\,(f_0(P)+f_3(P))?
168: \]
169: To study this question, we introduce the \emph{fatness} parameter
170: \[
171: \phi(P)\;\eqdef\;\frac{f_1(P)+f_2(P)}{f_0(P)+f_3(P)}
172: \]
173: of a $4$-polytope $P$.  We would like to know whether fatness is bounded.
174: 
175: For example, the $4$-simplex has fatness $2$, while the $4$-cube and the
176: $4$-cross polytope have fatness $\frac{56}{24}=\frac73$. More generally, if $P$
177: is simple, then we can substitute the Dehn-Sommerville relations
178: \[
179: f_2(P)=f_1(P)+f_3(P)-f_0(P)\qquad f_1(P)=2f_0(P)
180: \]
181: into the formula for fatness, yielding
182: \[
183: \phi(P)=\frac{f_1(P)+f_2(P)}{f_0(P)+f_3(P)}
184:     = \frac{3f_0(P)+f_3(P)}{f_0(P)+f_3(P)}<3.
185: \]
186: 
187: Since every $4$-polytope and its dual have the same fatness, the same upper
188: bound holds for simplicial $4$-polytopes. On the other hand, the ``neighborly
189: cubical'' $4$-polytopes of Joswig and Ziegler \cite{JZ:cubical} have
190: $f$-vectors
191: \[
192: (4,2n,3n-6,n-2)\cdot 2^{n-2},
193: \]
194: and thus fatness
195: \[
196: \phi=\frac{5n-6}{n+2} \to 5.
197: \]
198: In particular, the construction of these polytopes disproved the conjectured
199: flag-vector inequalities of Bayer \cite[pp.~145, 149]{Bayer:extended} and
200: Billera and Ehrenborg \cite[p.~109]{HZ:census}.
201: 
202: The main results of this paper are two lower bounds on fatness:
203: 
204: \begin{theorem} There are convex $4$-polytopes $P$ with fatness
205: $\phi(P)>\maxfatness$. \label{th:fat} \end{theorem}
206: 
207: \begin{theorem} The fatness of cellulated $3$-spheres is not bounded.
208: A $3$-sphere $S$ with $N$ vertices may have fatness as high as
209: $\phi(S) = \Omega(N^{1/12})$.
210: \label{th:fatter} \end{theorem}
211: 
212: We will prove \thm{th:fat} in \sec{s:poly} and \thm{th:fatter} in
213: \sec{s:spheres}, and present a number of related results along the
214: way.
215: 
216: \section{Conventions}
217: \label{s:conven}
218: 
219: Let $X$ be a finite
220: CW complex. If $X$ is identified with a manifold $M$, it is also called a
221: \emph{cellulation} of $M$. The complex $X$ is \emph{regular} if all closed
222: cells are embedded \cite[\S38]{Munkres:algtop}.
223: If $X$ is regular, we define it to
224: be \emph{strongly regular} if in addition the intersection of any
225: two closed cells
226: is a cell.  For example, every simplicial complex is a strongly regular CW
227: complex.  The complex $X$ is \emph{perfect} if the boundary maps of its chain
228: complex vanish.  (A non-zero-dimensional perfect complex is never regular.)
229: 
230: The \emph{$f$-vector} of a cellulation $X$, denoted $f(X)=(f_0,f_1,\ldots)$,
231: counts the number of cells in each dimension:  $f_0(X)$ is the number of
232: vertices, $f_1(X)$ is the number of edges, etc.
233: 
234: If $X$ is $2$-dimensional, we define its fatness as
235: \[
236: \phi(X) \eqdef \frac{f_1(X)}{f_0(X) + f_2(X)}.
237: \]
238: If $X$ is $3$-dimensional, we define its fatness $X$ as
239: \[
240: \phi(X) \eqdef \frac{f_1(X) + f_2(X)}{f_0(X) + f_3(X)}.
241: \]
242: If $P$ is a convex $d$-polytope, then its $f$-vector $f(P)$ is defined to be
243: the $f$-vector of its boundary complex, which is a strongly regular
244: $(d-1)$-sphere. If $d=4$ we can thus consider $\phi(P)$, the fatness of $P$.
245: The faces of $P$ of dimension 0 and 1 are called \emph{vertices} and
246: \emph{edges} while the faces of dimension $d-1$ and $d-2$ are called
247: \emph{facets} and \emph{ridges}. We extend this terminology to general
248: cellulations of $(d-1)$-manifolds.
249: 
250: The \emph{flag vector} of a regular cellulation $X$, and likewise the flag
251: vector of a polytope $P$, counts the number of nested sequences of cells with
252: prespecified dimensions.  For example, $f_{013}(X)$ is the number triples
253: consisting of a $3$-cell of $X$, an edge of the $3$-cell, and a vertex of the
254: edge; if $P$ is a 4-cube, then
255: $$f_{013}(P) = 192.$$
256: 
257: A convex polytope $P$ is \emph{simplicial} if each facet of $P$ is a simplex.
258: It is \emph{simple} if its polar dual $P\pol$
259: is simplicial, or equivalently if the cone of each vertex matches
260: that of a simplex.
261: 
262: 
263: \section{4-Polytopes}
264: \label{s:poly}
265: 
266: In this section we construct families of $4$-polytopes with several
267: interesting properties:
268: 
269: \begin{itemize}\setlength{\itemsep}{0pt}
270: \item They are the first known infinite families of {\em $2$-simple,
271: $2$-simplicial $4$-polytopes}, that is, polytopes in which all $2$-faces and
272: all dual $2$-faces are triangles (all edges are ``co-simple''). \Eg, Bayer
273: \cite{Bayer:extended} says that it would be interesting to have an infinite
274: family.
275: 
276: As far as we know, there were only six such polytopes previously known: the
277: simplex, the hypersimplex (the set of points in $[0,1]^4$ with coordinate sum
278: between $1$ and $2$), the dual of the hypersimplex, the $24$-cell, and a gluing
279: of two hypersimplices (due to Braden \cite{Braden:glued}), and the dual of the
280: gluing. (There are claims in Gr\"unbaum
281: \cite[p.~82, resp. p.~170]{Grunbaum:convex} that results of Shephard, resp.\
282: Perles and Shephard, imply the existence of infinitely many $2$-simple
283: $2$-simplicial $4$-polytopes.  Both claims appear to be incorrect.)
284: 
285: \item They are the fattest known convex $4$-polytopes.
286: 
287: \item They yield finite packings of (not necessarily congruent) spheres in
288: $\R^3$ with slightly higher average kissing numbers than previously known
289: examples \cite{Kuperberg:kissing}.
290: \end{itemize}
291: 
292: Let $Q\subset\R^4$ be a $4$-polytope that contains the origin in its interior.
293: If an edge $e$ of $Q$ is tangent to the unit sphere $S^3\subset\R^4$ at a point
294: $t$, then the corresponding ridge ($2$-dimensional face) $F=e\ppol$ of the
295: polar dual $P=Q\pol$ is also tangent to $S^3$ at $t$.  (Recall that the
296: polar dual is defined as
297: $$Q\pol \eqdef \{p | p \cdot q \le 1 \forall q \in Q\}.)$$
298: Furthermore, the affine
299: hulls of $e$ and $F$ form orthogonal complements in the tangent space of $S^3$,
300: so the convex hull $\conv(e\cup F)$ is an orthogonal bipyramid tangent to $S^3$
301: (\cf, Schulte \cite[Thm.~1]{Schulte:steinitz}).
302: 
303: We will construct polytopes $E$ by what we call the \emph{$E$-construction}.
304: This means that they are convex hulls
305: $$E\eqdef\conv(Q\cup P),$$
306: where $Q$ is a simplicial $4$-dimensional polytope whose edges are tangent to
307: the unit $3$-sphere $S^3$, and $P$ is the polar dual of $Q$.  Thus $P$ is
308: simple and its ridges are tangent to $S^3$. (We call $Q$ \emph{edge-tangent}
309: and $P$ \emph{ridge-tangent}.)
310: 
311: \begin{proposition}
312: If $P$ is a simple, ridge-tangent $4$-polytope, then the $4$-polytope
313: $E=\conv(P\cup Q)$ produced by the $E$-construction is $2$-simple and
314: $2$-simplicial, with $f$-vector
315: \[
316:  f(E) = (f_2(P), 6f_0(P), 6f_0(P), f_2(P)),
317: \]
318: and fatness
319: \[
320:  \phi(E) = 6\,\frac{f_0(P)}{f_2(P)}.
321: \]
322: \label{p:e-construction} \end{proposition}
323: 
324: \begin{proof}
325: Another way to view the $E$-construction is that $E$ is produced from $P$ by
326: adding the vertices of $Q$ sequentially. At each step, we cap a facet of $P$
327: with a pyramid whose apex is a vertex of $Q$. Thus the new facets consist of
328: pyramids over the ridges of $P$, where two pyramids with the same base
329: (appearing in different steps) lie in the same hyperplane (tangent to $S^3$),
330: and together form a bipyramid.  The facets of the final polytope $E$ are
331: orthogonal bipyramids over the ridges of $P$ and are tangent to $S^3$.  Since
332: the 2-faces of $E$ are pyramids over the edges of $P$, $E$ is $2$-simplicial.
333: 
334: The polytope $E$ is $2$-simple if and only if each edge is \emph{co-simple},
335: \ie, contained in exactly three facets of $E$.  The iterative construction of
336: $E$ shows that it has two types of edges: (i) edges of $P$, which are co-simple
337: in $E$ if and only if they are co-simple in $P$, and (ii) edges formed by
338: adding pyramids, which are co-simple if and only if the facets of $P$ are
339: simple.  Since $P$ is simple, its facets are simple and its edges are
340: co-simple, so $E$ is then $2$-simple.
341: 
342: This combinatorial description of $E$ yields an expression for the $f$-vector
343: of $E$ in terms of the flag vector
344: \cite{BJ:dehn,Bayer:extended,HZ:census} of $P$.  Since the facets
345: of $E$ are bipyramids over the ridges of $P$, the following identities hold:
346: \begin{align*}
347: f_3(E) &= f_2(P) & f_2(E) &= f_{13}(P) \\
348: f_1(E) &= f_1(P)+f_{03}(P) & f_0(E) &= f_0(P)+f_3(P).
349: \end{align*}
350: Since $P$ is simple,
351: \[
352: f_{03}(P)=4f_0(P) \qquad f_{13}(P)=3f_1(P).
353: \]
354: These identities together with Euler's equation and $f_1(P)=2f_0(P)$ imply the
355: proposition.
356: \end{proof}
357: 
358: The $f$-vector of $P$ also satisfies
359: \[
360: f_0(P)-f_1(P)+f_2(P)-f_3(P)=0 \qquad f_1(P)=2f_0(P),
361: \]
362: in the second case because $P$ is simple, so the fatness of $E$ can also be
363: written
364: \[
365: \phi(E) = 6\,\Big(1-\frac{f_3(P)}{f_2(P)}\Big) =
366: 6\,\Big(1-\frac{f_0(Q)}{f_1(Q)}\Big).
367: \]
368: Thus maximizing the fatness of $E$ is equivalent to maximizing the ridge-facet
369: ratio $f_2(P)/f_3(P)$, or the average degree of the graph of $Q$.  It also
370: shows that the $E$-construction cannot achieve a fatness of $6$ or more.
371: 
372: In light of \prop{p:e-construction}, we would like to construct edge-tangent
373: simplicial $4$-polytopes. Regular simplicial $4$-polytopes (suitably scaled)
374: provide three obvious examples: the $4$-simplex $\Delta_4$, the cross polytope
375: $C_4\pol$, and the $600$-cell.  From these, the $E$-construction produces the
376: dual of the hypersimplex, the 24-cell, and a new $2$-simple, $2$-simplicial
377: polytope with $f$-vector $(720,3600,3600,720)$ and fatness $5$,
378: whose facets are bipyramids over pentagons.
379: 
380: We will construct new edge-tangent simplicial $4$-polytopes by gluing together
381: (not necessarily simplicial) edge-tangent $4$-polytopes, called \emph{atoms},
382: to form \emph{compounds}.  We must position the polytopes so that their
383: facets match, they remain edge-tangent, and the resulting compound is convex.
384: It will be very useful to interpret the interior of the $4$-dimensional unit
385: ball as the Klein model of hyperbolic $4$-space $\HH^4$, with $S^3$ the sphere
386: at infinity.  (See Iversen \cite{Iverson:hyperbolic} and Thurston
387: \cite[Chap.~2]{Thurston:geometry} for introductions to hyperbolic geometry.) In
388: particular, Euclidean lines are straight in the Klein model, Euclidean
389: subspaces are flat, and hence any intersection of a convex polytope with
390: $\HH^4$ is a convex (hyperbolic) polyhedron.  Even though the Klein model
391: respects convexity, it does not respect angles. However, angles and convexity
392: are preserved under hyperbolic isometries.  There are enough isometries to
393: favorably position certain $4$-polytopes to produce convex compounds.
394: 
395: If a polytope $Q$ is edge-tangent to $S^3$, then it is hyperbolically
396: hyperideal:  Not only its vertices, but also its edges, lie beyond the sphere
397: at infinity, except for the tangency point of each edge.  Nonetheless portions
398: of its facets and ridges lie in the finite hyperbolic realm.  As a hyperbolic
399: object the polytope $Q$ (more precisely, $Q\cap \HH^4$) is convex and has flat
400: facets. The ridge $r$ between any two adjacent facets has a well-defined
401: hyperbolic dihedral angle, which is strictly between $0$ and $\pi$ if (as in
402: our situation) the ridge properly intersects $\HH^4$.  To compute this angle we
403: can intersect $r$ at any point $t$ with any hyperplane $R$ that contains the
404: (hyperbolic) orthogonal complement to $r$ at $t$.  We let $t$ be the tangent
405: point of any edge $e$ of the ridge, and let $R$ be the hyperplane perpendicular
406: to $e$.
407: 
408: \begin{fullfigure}{f:hyplink}{A cone emanating from an ideal point $t$ in
409:     $\HH^3$ in the Poincar\'e model, and a horosphere $S$ incident to $t$.  The
410:     link of $t$ (here a right isosceles triangle) inherits Euclidean geometry
411:     from $S$}
412: \begin{large} \pspicture(-5.5,-5.2)(5.2,6.5)
413: \pscircle(0,0){5}
414: \pscurve[linestyle=dashed](2.5,2.165)(1.294,2.415)(0,2.5)(-1.294,2.415)
415:     (-2.5,2.165)(-3.536,1.768)(-4.33,1.25)(-4.83,0.647)(-5,0)
416: \pscurve[linestyle=dashed](2.5,2.165)(3.536,1.768)(4.33,1.25)(4.83,0.647)(5,0)
417: \pscurve(-5,0)(-4.83,-0.647)(-4.33,-1.25)(-3.536,-1.768)(-2.5,-2.165)
418:     (-1.294,-2.415)(0,-2.5)(1.294,-2.415)(2.5,-2.165)(3.536,-1.768)(4.33,-1.25)
419:     (4.83,-0.647)(5,0)
420: \pscircle(1.5,1.299){2}
421: \pscurve[linestyle=dashed](2.5,2.165)(2.018,2.265)(1.5,2.299)(0.982,2.265)
422:     (0.5,2.165)(0.086,2.006)(-0.232,1.799)(-0.432,1.558)(-0.5,1.299)
423: \pscurve[linestyle=dashed](2.5,2.165)(3.032,1.942)(3.379,1.641)(3.5,1.299)
424: \pscurve(-0.5,1.299)(-0.379,0.957)(-0.032,0.656)(0.5,0.433)(1.153,0.314)
425:     (1.847,0.314)(2.5,0.433)(3.032,0.656)(3.379,0.957)(3.5,1.299)
426: \pscurve[linestyle=dashed](2.5,2.165)(2.121,1.887)(1.751,1.715)(1.398,1.654)
427:     (1.07,1.705)(0.776,1.866)
428: \pscurve(0,4.33)(0.033,3.66)(0.13,3.064)(0.289,2.559)(0.507,2.156)(0.776,1.866)
429: \pscurve[linestyle=dashed](2.5,2.165)(2.171,1.822)(1.95,1.459)(1.841,1.082)
430:     (1.848,0.702)(1.97,0.327)
431: \pscurve(4.33,-1.25)(3.665,-1.02)(3.082,-0.74)(2.597,-0.416)(2.224,-0.057)
432:     (1.97,0.327)
433: \psline[linestyle=dashed](2.5,2.165)(0.5,0.433)
434: \psline(0.5,0.433)(-2.5,-2.165)
435: \pscurve(-2.5,-2.165)(-2.415,-0.971)(-2.165,0.29)(-1.768,1.531)(-1.25,2.667)
436:     (-0.647,3.622)(-0,4.33)
437: \pscurve(4.33,-1.25)(4.183,-0.087)(3.75,1.083)(3.062,2.178)(2.165,3.125)
438:     (1.121,3.859)(0,4.33)
439: \pscurve[linewidth=1.5pt](0.5,0.433)(0.512,0.706)(0.546,0.992)(0.602,1.285)
440:     (0.68,1.579)(0.776,1.866)
441: \pscurve[linewidth=1.5pt](0.5,0.433)(0.774,0.367)(1.065,0.323)(1.366,0.301)
442:     (1.67,0.303)(1.97,0.327)
443: \pscurve[linewidth=1.5pt](0.776,1.866)(0.973,1.581)(1.22,1.244)(1.491,0.894)
444:     (1.751,0.578)(1.97,0.327)
445: \qdisk(4.33,-1.25){1.8pt}
446: \qdisk(-2.5,-2.165){1.8pt}
447: \qdisk(2.5,2.165){1.8pt}
448: \qdisk(0,4.33){1.8pt}
449: \pspolygon[linewidth=1.5pt](-5.3,5)(-5.3,6.3)(-4,5)
450: \pcline[nodesepA=1.3,nodesepB=1,arrows=->](-4.867,5.433)(1.082,0.875)
451: \uput[-75](2.5,2.165){$t$}
452: \rput(1.5,-0.26){$S$}
453: \rput(-3.348,0.1){$\HH^3$}
454: \endpspicture
455: \end{large} \eatline \end{fullfigure}
456: 
457: Within the hyperbolic geometry of $R \cong \HH^3$, every line emanating from
458: the ideal point $t$ is orthogonal to any horosphere incident to $t$. Thus the
459: link of the edge $e$ of $Q$ is the intersection of $Q\cap R$ with a
460: sufficiently small horosphere $S$ at $t$.  Since horospheres have flat
461: Euclidean geometry \cite[p.~61]{Thurston:geometry}, the link $Q \cap S$ is a
462: Euclidean polygon.  Its edges correspond to the facets of $Q$ that contain $e$,
463: and its vertices correspond to the ridges of $Q$ that contain $e$. Thus the
464: dihedral angle of a ridge $r$ of $Q$ equals to the Euclidean angle of the
465: vertex $r \cap S$ of the Euclidean polygon $Q \cap S$. This is easier to see in
466: the Poincar\'e model of hyperbolic space, because it respects angles, than in
467: the Klein model. \fig{f:hyplink} shows an example.
468: 
469: To summarize:
470: 
471: \begin{lemma}
472: A compound of two or more polytopes is convex if and only if each ridge has
473: hyperbolic dihedral angle less than $\pi$, or equivalently, iff each edge
474: link is a convex Euclidean polygon.
475: \label{l:convex} \end{lemma}
476: 
477: Compounds can also have interior ridges with total dihedral
478: angle exactly $2\pi$.  But since all atoms of a compound are
479: edge-tangent, compounds do not have any interior edges or vertices.
480: 
481: If $Q$ is a regular polytope, then $Q \cap S$ is a regular polygon.
482: The following lemma is then immediate:
483: 
484: \begin{lemma}
485: If $Q$ is a regular, edge-tangent, simplicial $4$-polytope,
486: then in the hyperbolic metric of the Klein model, its dihedral angles
487: are $\pi/3$ (for the simplex), $\pi/2$ (for the cross polytope), and
488: $3\pi/5$ (for the $600$-cell).
489: \label{l:regang} \end{lemma}
490: 
491: A hyperideal hyperbolic object, even if it is an edge-tangent convex
492: polytope, can be unfavorably positioned so that it is unbounded as a Euclidean
493: object (\cf, Schulte \cite[p.~508]{Schulte:steinitz}).  Fortunately
494: there is always a bounded position as well:
495: 
496: \begin{lemma} Let be $Q$ an edge-tangent, convex polytope in $\R^d$ whose
497: points of tangency with $S^{d-1}$ do not lie in a hyperplane. Then there is a
498: hyperbolic isometry $h$ (extended to all of $\R^d$) such that $h(Q)$ is
499: bounded.
500: \label{l:bounded} \end{lemma}
501: \begin{proof} Let $p$ lie in the interior of the convex hull of the edge
502: tangencies and let $f$ be any hyperbolic motion that moves $p$ to the Euclidean
503: origin in $\R^d$.  Since the convex hull $K$ of the edge tangencies of $h(Q)$
504: contains the origin, $K\pol$ is a bounded polytope that circumscribes
505: $S^{d-1}$. Since $K\pol$ is facet-tangent where $h(Q)$ is edge-tangent, $h(Q)
506: \subset K\pol$.
507: \end{proof}
508: 
509: In the following we discuss three classes of edge-tangent simplicial convex
510: $4$-polytopes that are obtained by gluing in the Klein model: Compounds of
511: simplices, then simplices and cross polytopes, and finally compounds from cut
512: $600$-cells.  There are yet other edge-tangent compounds involving cross
513: polytopes cut in half (\ie, pyramids over octahedra), $24$-cells, and
514: hypersimplices as atoms, but we will not discuss these here.
515: 
516: \subsection{Compounds of simplices}
517: 
518: In this section we classify compounds whose atoms are simplices. This includes
519: all \emph{stacked} polytopes, which are simplicial polytopes that decompose as
520: a union of simplices without any interior faces other than facets. However
521: compounds of simplices are a larger class, since they may have interior ridges.
522: 
523: \begin{lemma}
524: Any edge-tangent $d$-simplex is hyperbolically regular.
525: \label{l:hypreg} \end{lemma}
526: 
527: \begin{proof} The proof is by induction on $d$, starting from the case $d=2$, where the
528: three tangency points define an ideal triangle in $\HH^2$. All ideal triangles
529: are congruent \cite[p.~83]{Thurston:geometry}.  Since edge-tangent triangles
530: are the polar duals of ideal triangles, they are all equivalent as well. 
531: 
532: If $d>2$, let $B$ be a general edge-tangent $d$-simplex. On the one hand, there
533: exists an edge-tangent simplex $A$ which is regular both in hyperbolic geometry
534: and Euclidean geometry.   On the other hand, given the position of $d$ of the 
535: vertices, there are at most two choices for the last vertex that  produce an
536: edge-tangent simplex, one on each side of the hyperplane spanned by the first
537: $d$. By induction there exists an isometry that takes a face of $B$ to a face
538: of $A$ and the remaining vertex to the same side.  The edge-tangent constraint
539: implies that this isometry takes the last vertex of $B$ to the last vertex of
540: $A$ as well.
541: \end{proof}
542: 
543: \begin{proposition}\label{p:simplices}
544: There are only three possible edge-tangent compounds of $4$-simplices:
545: \begin{itemize}\setlength{\itemsep}{0pt}
546: \item the regular simplex,
547: \item the bipyramid (a compound of two simplices that share a facet), and
548: \item the join of a triangle and a hexagon (a compound of six
549:   simplices that share a ridge).
550: \end{itemize}
551: \end{proposition}
552: 
553: \begin{proof} \fig{f:tjewels} shows all strictly convex polygons with
554: unit-length edges tiled by unit equilateral triangles, or \emph{triangle
555: jewels}.  Since the atoms of an edge-tangent compound of simplices are
556: edge-tangent, they are hyperbolically regular by \lem{l:hypreg}, and their edge
557: links are equilateral triangles. Thus every edge link of a compound of
558: simplices is a triangle jewel.  Any three $4$-simplices in a chain in such a
559: compound share a ridge. In order to create an edge link matching
560: \fig{f:tjewels}, they must extend to a ring of six simplices around the same
561: ridge. Adding any further simplex to these six would create an edge link in the
562: form of a triangle surrounded by three other triangles, which does not appear
563: in \fig{f:tjewels}.
564: \end{proof}
565: 
566: \begin{fullfigure}{f:tjewels}{The 3 possible edge links of edge-tangent
567:     compounds of $4$-simplices}
568: \psset{unit=.8cm} \pspicture(-2.5,-1)(3.5,1)
569: \rput(-1.5,0){\pspicture(-.5,-.6)(.5,.6) % triangle
570: \pspolygon(0,.577)(-.5,-.289)(.5,-.289)
571: \endpspicture}
572: \rput(0,0){\pspicture(-.5,-1)(.5,1) % lozenge
573: \psline(-.5,0)(0,.866)(.5,0)(-.5,0)(0,-.866)(.5,0)
574: \endpspicture}
575: \rput(2,0){\pspicture(-1,-1)(1,1) % hexagon
576: \pspolygon(1;0)(1;60)(1;120)(1;180)(1;240)(1;300)
577: \psline(1;0)(1;180) \psline(1;60)(1;240) \psline(1;120)(1;300)
578: \endpspicture}
579: \endpspicture \eatline \end{fullfigure}
580: 
581: Two of the $E$-polytopes produced by \prop{p:simplices} were previously known.
582: If $Q$ is the simplex, then $E$ is dual to the hypersimplex.  If $Q$ is the
583: bipyramid, then $E$ is dual to Braden's glued hypersimplex.  However, if $Q$ is
584: the six-simplex compound (dual to the product of a triangle and a hexagon),
585: then $E$ is a new $2$-simple, $2$-simplicial polytope with $f$-vector
586: $(27,108,108,27)$.
587: 
588: \prop{p:simplices} also implies an interesting impossibility
589: result.
590: 
591: \begin{corollary}
592: No stacked $4$-polytope with more than $6$ vertices is edge-tangent.
593: \end{corollary}
594: 
595: See Schulte~\cite[Thm.~3]{Schulte:steinitz} for the first examples of
596: polytopes that have no edge-tangent realization.
597: 
598: \subsection{Compounds of simplices and cross polytopes}
599: 
600: Next, we consider compounds of simplices and regular cross polytopes. The edge
601: link of any convex compound of these two types of polytopes must be one of the
602: eleven strictly convex polygons tiled by unit triangles and squares, or
603: \emph{square-triangle jewels} (\fig{f:stjewels}).  See Malkevitch
604: \cite{Malkevitch:tiling} and Waite \cite{Waite:combining} for work on convex
605: compounds of these shapes relaxing the requirement of strict convexity.
606: 
607: \begin{fullfigure}{f:stjewels}{The 11 possible edge links of edge-tangent
608:     compounds of $4$-simplices and cross polytopes}
609: \psset{unit=.8cm} \pspicture(-.75,.75)(10,8.75)
610: \rput(0,8){\pspicture(-.5,-.5)(.5,.5) % triangle
611: \pspolygon(0,.433)(-.5,-.433)(.5,-.433) \endpspicture}
612: \rput(1.5,8){\pspicture(-.5,-.5)(.5,.5) % square
613: \psframe(-.5,-.5)(.5,.5) \endpspicture}
614: \rput(3.25,8){\pspicture(-1,-.5)(1,.5) % lozenge
615: \psline(0,-.5)(.866,0)(0,.5)(0,-.5)(-.866,0)(0,.5) \endpspicture}
616: \rput(5.5,8){\pspicture(-.5,-.5)(1.5,.5) % house
617: \psframe(-.5,-.5)(.5,.5) \psline(.5,.5)(1.366,0)(.5,-.5)
618: \endpspicture}
619: \rput(8.25,8){\pspicture(-1.5,-.5)(1.5,.5) % football
620: \psframe(-.5,-.5)(.5,.5) \psline(.5,.5)(1.366,0)(.5,-.5)
621: \psline(-.5,.5)(-1.366,0)(-.5,-.5) \endpspicture}
622: \rput(.75,6.25){\pspicture(-1,-1)(1,1) % hexagon
623: \pspolygon(1;0)(1;60)(1;120)(1;180)(1;240)(1;300)
624: \psline(1;0)(1;180) \psline(1;60)(1;240) \psline(1;120)(1;300)
625: \endpspicture}
626: \rput(4,6){\pspicture(-1.5,-1)(1.5,1) % six+two
627: \pspolygon(.5,1)(-.5,1)(-1.366,.5)(-1.366,-.5)(-.5,-1)
628:     (.5,-1)(1.366,-.5)(1.366,.5)
629: \psline(-.5,-1)(-.5,1) \psline(.5,-1)(.5,1)
630: \psline(1.366,.5)(.5,0)(-.5,0)(-1.366,.5)
631: \psline(1.366,-.5)(.5,0) \psline(-.5,0)(-1.366,-.5)
632: \endpspicture}
633: \rput(7.75,5.75){\pspicture(-1.5,-1)(1.5,1) % eight+four
634: \pspolygon(.5,1.366)(-.5,1.366)(-1.366,.866)(-1.866,0)(-1.366,-.866)
635:     (-.5,-1.366)(.5,-1.366)(1.366,-.866)(1.866,0)(1.366,.866)
636: \pspolygon(0,-.5)(.866,0)(0,.5)(-.866,0)
637: \psline(1.866,0)(.866,0) \psline(-1.866,0)(-.866,0)
638: \psline(1.366,.866)(.866,0)(1.366,-.866)
639: \psline(-1.366,.866)(-.866,0)(-1.366,-.866)
640: \psline(.5,1.366)(0,.5)(-.5,1.366) \psline(.5,-1.366)(0,-.5)(-.5,-1.366)
641: \psline(0,.5)(0,-.5)
642: \endpspicture}
643: \rput(1.5,3){\pspicture(-2,-2)(2,2) % pie
644: \pspolygon(1.932;15)(1.932;45)(1.932;75)(1.932;105)(1.932;135)(1.932;165)
645:     (1.932;195)(1.932;225)(1.932;255)(1.932;285)(1.932;315)(1.932;345)
646: \pspolygon(1;30)(1;90)(1;150)(1;210)(1;270)(1;330)
647: \psline(1.932;15)(1;30)(1.932;45) \psline(1.932;75)(1;90)(1.932;105)
648: \psline(1.932;135)(1;150)(1.932;165) \psline(1.932;195)(1;210)(1.932;225)
649: \psline(1.932;255)(1;270)(1.932;285) \psline(1.932;315)(1;330)(1.932;345)
650: \psline(1;30)(1;210) \psline(1;90)(1;270) \psline(1;150)(1;330)
651: \endpspicture}
652: \rput(5.25,3.5){\pspicture(-2,-2)(2,2) % seven+three
653: \pspolygon(0,-1.866)(.866,-1.366)(1.366,-.5)(1.366,.5)(.5,1)(-.5,1)
654:     (-1.366,.5)(-1.366,-.5)(-.866,-1.366)
655: \pspolygon(0,-.866)(.5,0)(-.5,0)
656: \psline(1.366,-.5)(.5,0)(.5,1) \psline(-1.366,-.5)(-.5,0)(-.5,1)
657: \psline(-.866,-1.366)(0,-.866)(.866,-1.366)
658: \psline(1.366,.5)(.5,0) \psline(-1.366,.5)(-.5,0) \psline(0,-1.866)(0,-.866)
659: \endpspicture}
660: \rput(8.5,3){\pspicture(-2,-1.5)(0,2) % three+two
661: \pspolygon(-0,.5)(-.5,1.366)(-1.366,.866)(-1.866,0)(-1.366,-.866)
662:     (-.5,-1.366)(-0,-.5)
663: \psline(-.866,0)(-0,.5) \psline(-.866,0)(0,-.5) \psline(-.866,0)(-1.366,.866)
664: \psline(-.866,0)(-1.866,0) \psline(-.866,0)(-1.366,-.866)
665: \endpspicture}
666: \endpspicture \eatline \end{fullfigure}
667: 
668: If $Q$ is a single cross polytope, then $E$ is a 24-cell. We can also glue
669: simplices onto subsets of the facets of the cross polytope. The new dihedral
670: angles formed by such a gluing are $5\pi/6$. The resulting compound is convex
671: as long as no two glued cross polytope facets share a ridge.  We used a
672: computer program to list the combinatorially distinct ways of choosing a subset
673: of nonadjacent facets of the cross polytope; the results may be summarized as
674: follows. In addition to the 24-cell, this yields 20 new $2$-simple,
675: $2$-simplicial polytopes.
676: 
677: \begin{proposition}
678: There are exactly $21$ distinct simplicial edge-tangent compounds composed of
679: one regular $4$-dimensional cross polytope and $k\ge0$ simplices, according to
680: the following table:
681: $$\begin{tabular}{r@{\;}|@{\;}ccccccccc@{\;}|@{\;}c}
682: $k$&0&1&2&3&4&5&6&7&8&Total\\ \hline
683: \#&1&1&3&3&6&3&2&1&1&21\\
684: \end{tabular}$$
685: \label{p:crosspoly} \end{proposition}
686: 
687: We can also confirm that every square-triangle jewel arises as the edge link of
688: an edge-tangent $4$-dimensional compound.  For every jewel other than the one
689: in the center, we can form a convex \emph{edge bouquet} consisting of simplices
690: and cross polytopes that meet at an edge: we replace each triangle by a simplex
691: and each square by a cross polytope.  Since the central jewel has two adjacent
692: squares, its edge bouquet is not convex.  Instead we glue two cross polytopes
693: along a facet so that the 4 ridges of that facet are flush, \ie, their dihedral
694: angle is $\pi$.  Thus we can ``caulk'' each such ridge with three simplices
695: that share the ridge.  The central jewel is the link of 6 of the edges of the
696: resulting compound of 2 cross polytopes and 12 simplices.
697: 
698: Simplices and regular cross polytopes combine to form many other edge-tangent
699: simplicial polytopes and hence $2$-simple, $2$-simplicial
700: polytopes. In particular, these methods lead to the following theorem.
701: 
702: \begin{theorem}
703: There are infinitely many combinatorially distinct $2$-simple, $2$-simplicial
704: facet-tangent $4$-polytopes.
705: \label{th:inf} \end{theorem}
706: 
707: \begin{proof} We glue $n$ cross polytopes end-to-end. Each adjacent pair
708: produces 4 flush ridges that we caulk with chains of three simplices.  The
709: facets to which these simplices are glued are not adjacent and so do not
710: produce any further concavities.
711: \end{proof}
712: 
713: The chain of $n$ cross polytopes has $f$-vector
714: \[
715: (4n+4,18n+6,28n+4,14n+2).
716: \]
717: Filling a concavity adds $(2,9,14,7)$, so after filling the $4(n-1)$
718: concavities we get a simplicial polytope $Q$ with
719: \[
720: f(Q)\;=\;(12n- 4,54n-30,84n-54,42-26),
721: \]
722: which yields a $2$-simple, $2$-simplicial $4$-polytope $E$
723: with
724: \[
725: f(E)\;=\;(54n-30,252n-156,252n-156,54n-30)
726: \]
727: by \prop{p:e-construction}.
728: 
729: \begin{remark} Every $2$-simple, $2$-simplicial $4$-polytopes that we know is
730: combinatorially equivalent to one which circumscribes the sphere.  Are there
731: any that are not?
732: 
733: We do know a few $2$-simple, $2$-simplicial $4$-polytopes which are not
734: $E$-polytopes.  Trivially there is the simplex. There are a few others that
735: arise by the fact that the $24$-cell is the $E$-polytope of a cross polytope
736: in 3 different ways.  Color the vertices of a $24$-cell red, green, and blue,
737: so that the vertices of each color span a cross polytope. If we cap one facet
738: of a cross-polytope by a simplex and apply the $E$-construction, the result is
739: a $24$-cell in which 6 facets that meet at 1 vertex are replaced by 10 facets
740: and 4 vertices.  If the replaced vertex is red, the replacement can be induced
741: by capping either blue cross polytope or the green cross polytope and then
742: applying the $E$-construction; the position of the replacement differs between
743: the two cases. If we replace two different red vertices, one by capping the
744: green cross polytope and the other by capping the blue cross polytope, then
745: the resulting polytope is $2$-simple and $2$-simplicial but not an
746: $E$-polytope.  This construction has several variations: for example, we can
747: also replace three vertices, one of each color.
748: \end{remark}
749: 
750: \subsection{Compounds involving the $600$-cell}
751: \label{s:600}
752: 
753: If $Q$ is the $600$-cell, then $E$ is a $2$-simple, $2$-simplicial polytope
754: with $f$-vector $(720,3600,3600,720)$ and fatness exactly $5$.  Again, we can
755: glue simplices onto any subset of nonadjacent facets of the $600$-cell,
756: creating convex compounds with dihedral angle $14\pi/15$. We did not count the
757: (large) number of distinct ways of choosing such a subset, analogous to
758: \prop{p:crosspoly}. It is not possible to glue a cross polytope to a
759: $600$-cell, because that would create an $11\pi/10$ angle (\ie, a concave
760: dihedral of $9\pi/10$) which cannot be filled by additional simplices or cross
761: polytopes.
762: 
763: The large dihedral angles of the $600$-cell make it difficult to form
764: compounds from it, but we can modify it as follows to create smaller dihedrals.
765: Remove a vertex and form the convex hull of the remaining 119 vertices.
766: The resulting convex polytope has 580 of the $600$-cell's tetrahedral facets
767: and one icosahedral facet. The pentagonal edge link (\fig{f:cutpents}(a)) of
768: the edges bordering this new facet become modified in a similar way, by
769: removing one vertex and forming the convex hull of the remaining four vertices
770: (\fig{f:cutpents}(b)), which results in a trapezoid; thus, the hyperbolic
771: dihedrals at the ridges around the new facet are $2\pi/5$.
772: 
773: \begin{fullfigure}{f:cutpents}{Edge figures of (a) a $600$-cell, (b) a
774:     $600$-cell with one vertex removed, (c) a $600$-cell with two removed
775:     vertices, and (d) an icosahedral cap}
776: \pspicture(0,-1)(9,10)
777: \rput(2,8){\pspicture(-2,-2)(2,2)
778: \pspolygon(2;90)(2;162)(2;234)(2;306)(2;18)
779: \uput[270](2;90){$\frac{3\pi}{5}$} \psarc(2;90){1.1}{216}{324}
780: \endpspicture} \rput(2,5.5){\large (a)}
781: \rput(7,8){\pspicture(-2,-2)(1.25,2)
782: \pspolygon(2;90)(2;162)(2;234)(2;306)
783: \uput{.4}[253](2;90){$\frac{2\pi}{5}$} \psarc(2;90){1.25}{216}{288}
784: \endpspicture} \rput(7,5.5){\large (b)}
785: \rput(2,2){\pspicture(-1.25,-2)(1.25,2)
786: \pspolygon(2;90)(2;234)(2;306)
787: \uput{.65}[270](2;90){$\frac{\pi}{5}$} \psarc(2;90){1.5}{252}{288}
788: \endpspicture} \rput(2,-.5){\large (c)}
789: \rput(7,2){\pspicture(-2,-2)(.25,2)
790: \pspolygon(2;90)(2;162)(2;234)
791: \uput{.65}[234](2;90){$\frac{\pi}{5}$} \psarc(2;90){1.5}{216}{252}
792: \endpspicture} \rput(7,-.5){\large (d)}
793: \endpspicture \eatline \end{fullfigure}
794: 
795: This cut polytope is not simplicial, but we may glue two of these polytopes
796: together along their icosahedral facets, forming a simplicial polytope with
797: $4\pi/5$ dihedrals along the glued ridges.  This compound's new edge links are
798: hexagons formed by gluing pairs of trapezoids (\fig{f:link600}(a)).  The
799: same cutting and gluing process may be repeated to form a sequence or tree of
800: $600$-cells, connected along cuts that do not share a ridge.  For such a chain
801: or tree formed from $n$ cut $600$-cells, the $f$-vector may be computed as
802: \[
803: f(Q)\;=\;(106n+14,666n+54,1120n+80,560n+40),
804: \]
805: so the $E$-construction yields
806: \[
807: f(E)\;=\;(666n+54,3360n+240,3360n+240,666n+54)
808: \]
809: and thus a fatness of
810: \[
811: \phi(E)\;=\;\frac{3360n+240}{666n+54}\ \
812: \longrightarrow\ \ \frac{560}{111}\ \approx\ 5.045045.
813: \]
814: Thus the fatness of the $2$-simple, $2$-simplicial polytopes formed by such
815: compounds improves slightly on that formed from the $600$-cell alone.
816: 
817: \begin{fullfigure}{f:link600}{Edge links of compounds of cut $600$-cells}
818: \pspicture(0,-5)(13.5,4)
819: \rput(2.5,0){\pspicture(-2,-2.5)(2,2.5)
820: \pspolygon(1.176,2.236)(-1.176,2.236)(-1.902,0)
821:     (-1.176,-2.236)(1.176,-2.236)(1.902,0)
822: \psline(-1.902,0)(1.902,0)
823: \endpspicture}
824: \rput(2.5,-4.5){\large (a)}
825: \rput(9.5,0){\pspicture(-4,-4)(4,4)
826: \pspolygon(3.804;0)(3.804;36)(3.804;72)(3.804;108)(3.804;144)
827:     (3.804;180)(3.804;216)(3.804;252)(3.804;288)(3.804;324)
828: \psline(3.804;0)(3.804;180) \psline(3.804;36)(3.804;216)
829: \psline(3.804;72)(3.804;252) \psline(3.804;108)(3.804;288)
830: \psline(3.804;144)(3.804;324)
831: \endpspicture}
832: \rput(9.5,-4.5){\large (b)}
833: \endpspicture
834: \end{fullfigure}
835: 
836: It is also possible to form compounds involving $600$-cells which have been cut
837: by removing several vertices (as described above) so that two of the resulting
838: icosahedral facets meet at a ridge. Each edge link at this ridge is an
839: isosceles triangle formed by removing two vertices from a pentagon
840: (\fig{f:cutpents}(c)). Thus the dihedral angle of the triangular ridge between
841: the icosahedral facets is $\pi/5$. We can therefore form compounds in which ten
842: of these doubly-cut $600$-cells share a triangle, whose edges links are a
843: regular decagon cut into ten isosceles triangles (\fig{f:link600}(b)).
844: Yet other compounds of cut $600$-cells and simplices are possible,
845: although we do not need them here.
846: 
847: The cut $600$-cells also form more complicated compounds which require some
848: group-theoretic terminology to explain.  The vertices of a regular $600$-cell
849: form a 120-element group under (rescaled) quaternionic multiplication, the
850: binary icosahedral group.  This group has a 24-element subgroup, the binary
851: tetrahedral group, which also arises as the units of the Hurwitz integers (see
852: Conway and Sloane \cite[\S2.2.6,8.2.1]{CS:splag}).  Let $A$ be the convex hull
853: of the other 96 vertices of the $600$-cell;  \ie, $A$ is formed by cutting 24
854: vertices from $600$-cell in the above manner. The resulting polytope is the
855: ``snub$\{3,4,3\}$'' (snub 24-cell) of Coxeter
856: \cite[\S8.4,8.5]{Coxeter:regpoly}.  Its $f$-vector is $(96,432,480,144)$.
857: Every icosahedral facet of $A$ is adjacent to $8$ other icosahedron facets, as
858: well as to $12$ tetrahedra. Thus $A$ has $96$ icosahedron-icosahedron ridges.
859: 
860: We can build new hyperbolic, edge-tangent, simplicial polytopes by gluing
861: copies of $A$ along icosahedral faces and capping the remaining icosahedral
862: facets with pyramidal caps of the type that we had cut off to form $A$. (The
863: edge links of such a cap $C$ are given by \fig{f:cutpents}(d).) The resulting
864: polytope $Q$ will be convex if at each icosahedral-icosahedral ridge of a copy
865: of $A$, either $10$ copies of $A$ meet, or two caps and one or two copies of
866: $A$ do.  Also at each icosahedral-tetrahedral ridge of a copy of $A$, either
867: two copies of $A$ or one each of $A$ and a cap must meet. If two copies of $A$
868: meet (in an icosahedral facet $F$), then they differ by a reflection through
869: $F$. These reflections generate a discrete hyperbolic reflection group $\Gamma$
870: since the supporting hyperplanes of the icosahedral facets (the facets of a
871: hyperideal $24$-cell, whose ridges are also ridges of $A$!) satisfy the Coxeter
872: condition: When they meet, they meet at an angle of $\pi/5$, which divides
873: $\pi$. Thus the copies of $A$ used in $Q$ are a finite subset $\Sigma$ of the
874: orbit of $A$ under $\Gamma$.  The set $\Sigma$ determines $Q$.
875: 
876: \begin{remark} That it suffices to consider the dihedral angles of adjacent
877: facets follows from Poincar\'e's covering-space argument:  Let $P$ be a
878: spherical, Euclidean, or hyperbolic polytope whose dihedral angles divide
879: $\pi$.  Let $X$ denote the space in which it lives.  Let $Y$ be the disjoint
880: union of all copies of $P$ in $X$ in every position, and let $Z$ be the
881: quotient of $Y$ given by identifying two copies of $P$ along a shared facet.
882: The space $Z$ is constructed abstractly so that $P$ tiles it.
883: 
884: We claim that $Z$ is a covering space of $X$.  Each $p \in Z$ lies in the
885: interior of some face $F$ of a copy of $P$.  If $F$ is a copy of $P$ or a
886: facet, this is elementary; if $F$ is a ridge, it follows from the dihedral
887: angle condition.  Otherwise it follows by applying the covering-space argument
888: inductively, replacing $X$ by the link $S$ of $F$ and $P$ by $P \cap S$.
889: 
890: Since $Z$ is a covering space, a connected component of $Z$ is a tiling of $X$
891: by $P$. See Vinberg \cite{Vinberg:reflection} for a survey of hyperbolic
892: reflection groups.
893: \end{remark}
894: 
895: There is no one best choice for $\Sigma$, only a supremal limit.  One
896: reasonable choice for $\Sigma$ is the corona of a copy of $A$, \ie, $A$
897: together with the set of all images under $\Gamma$ that meet it, necessarily at
898: a facet or a ridge. The corona of $A$ is depicted by a simplified (and
899: therefore erroneous) schematic in \fig{f:corona}; the reader should imagine the
900: correct, more complicated version.  The schematic uses a chemistry notation in
901: which each copy of $A$ is represented as an \emph{atom}, each pair of copies
902: that shares a facet is represented as a \emph{bond}, and ``$7A$'' denotes a
903: chain of 7 atoms.  To extend the terminology, we call 10 copies of $A$ that
904: meet at a ridge a \emph{ring}. The schematic is simplified in that the central
905: copy actually has 24 bonds (not 6), each of the neighbors has 9 bonds (not 3),
906: and there are 96 rings (not 6).
907: 
908: \begin{fullfigure}{f:corona}{An oversimplified schematic of a corona of $A$}
909: \begin{large} \pspicture(-4,-4)(4,4) \psset{nodesep=3pt}
910: \rput(0,0){\rnode{a}{$A$}}
911: \rput(2.4;0){\rnode{b1}{$A$}} \rput(2.4;60){\rnode{b2}{$A$}}
912: \rput(2.4;120){\rnode{b3}{$A$}} \rput(2.4;180){\rnode{b4}{$A$}}
913: \rput(2.4;240){\rnode{b5}{$A$}} \rput(2.4;300){\rnode{b6}{$A$}}
914: \rput(3.6;30){\rnode{c1}{$7A$}} \rput(3.6;90){\rnode{c2}{$7A$}}
915: \rput(3.6;150){\rnode{c3}{$7A$}} \rput(3.6;210){\rnode{c4}{$7A$}}
916: \rput(3.6;270){\rnode{c5}{$7A$}} \rput(3.6;330){\rnode{c6}{$7A$}}
917: \ncline{a}{b1} \ncline{a}{b2} \ncline{a}{b3}
918: \ncline{a}{b4} \ncline{a}{b5} \ncline{a}{b6}
919: \ncline{b1}{c1} \ncline{b1}{c6} \ncline{b2}{c2} \ncline{b2}{c1}
920: \ncline{b3}{c3} \ncline{b3}{c2} \ncline{b4}{c4} \ncline{b4}{c3}
921: \ncline{b5}{c5} \ncline{b5}{c4} \ncline{b6}{c6} \ncline{b6}{c5}
922: \endpspicture \end{large} \eatline \end{fullfigure}
923: 
924: Let $Q$ be the union of these copies of $A$ with the remaining icosahedral
925: facets capped.  To compute the $f$-vector of $Q$ it is easier to view each atom
926: as a copy of a 600-cell $B$, minus two caps for each bond.  There are $1+24+96
927: \cdot 7 = 697$ atoms and $24+ 96 \cdot 8 = 792$ bonds in total. Thus $Q$ has
928: \[
929: f_3(Q) = 697 f_3(B) - 792 \cdot 2 f_3'(C) = 386520
930: \]
931: facets, where $f_3'(C)=30$ is the number of simplicial facets of a cap $C$.
932: 
933: Counting vertices is more complicated.  Let $I$ be an icosahedron
934: and let $T$ be a triangle.  Then
935: \begin{align*}
936: f_0(Q) &= 697 f_0(B) - 792 \cdot 2 f_0(C) + 792 f_0(I) + 96 f_0(T) \\
937: &= 72840,
938: \end{align*}
939: because after the vertices of the caps are subtracted, the vertices of each
940: icosahedral facet at a bond are undercounted once, and after these are restored
941: the vertices of each triangle at the center of a ring are undercounted once.
942: The rest of the $f$-vector of $Q$ follows from the Dehn-Sommerville equations:
943: \begin{align*}
944: f_2(Q) &= 2f_3(Q) = 773040, \\
945: f_1(Q) &= f_0(Q) + f_3(Q) = 459360.
946: \end{align*}
947: The polytope $Q$ yields an $E$-polytope with fatness
948: \[
949: \phi(E) = 6\,\frac{f_3(Q)}{f_1(Q)} = \frac{3221}{638} \approx 5.048589.
950: \]
951: Note that since this bound arises from a specific choice of $\Sigma$
952: rather than a supremal limit, this is not optimal as a lower
953: bound of supremal fatness.
954: 
955: \subsection{Kissing numbers}
956: \label{s:kissing}
957: 
958: As mentioned above, another use of ridge-tangent polytopes $P$ is the average
959: kissing number problem \cite{Kuperberg:kissing}.  Let $X$ be a finite packing
960: of (not necessarily congruent) spheres $S^3$, which is equivalent to a finite
961: sphere packing in $\R^3$ by stereographic projection. The question is to
962: maximize the average number of kissing points of the spheres in $X$.  If $P$ is
963: ridge-tangent, its facets intersect the unit sphere $S^3$ in a sphere packing
964: $X$, in which the spheres kiss at the tangency points of $P$. Thus the
965: ridge-facet ratio of $P$ is exactly half the average kissing number of $X$.
966: (Not all sphere packings come from ridge-tangent polytopes in this way.)
967: 
968: The sphere packings due to Kuperberg and Schramm \cite{Kuperberg:kissing} can
969: be viewed as coming from a compound consisting of a chain or tree of $n$ cut
970: $600$-cells (\ie, atoms in the sense of \fig{f:corona}).  Their average kissing
971: numbers are
972: \[
973: \kappa\; =\; 2\,\frac{666n+54}{106n+14}\ \
974: \longrightarrow\ \ \frac{666}{53}\ \approx\ 12.56603.
975: \]
976: By contrast if $Q$ is the compound formed from a corona of $A$ as in
977: \sec{s:600}, then the average kissing number of the corresponding sphere
978: packing is
979: \[
980: \kappa(Q) = 2\,\frac{f_1(Q)}{f_0(Q)} = \frac{7656}{607} \approx 12.61285.
981: \]
982: Like the bound on fatness, it is not optimal as a lower bound
983: on the supremal average kissing number.
984: 
985: Here we offer no improvement on the upper bound
986: \[
987: \kappa < 8+4\sqrt{3} \approx 14.92820
988: \]
989: from \cite{Kuperberg:kissing}, even though it cannot be optimal either.
990: 
991: \section{3-Spheres}
992: \label{s:spheres}
993: 
994: In this section we construct a family of strongly regular cellulations of the
995: $3$-sphere with unbounded fatness.  Indeed, we provide an efficient version of
996: the construction, in the sense that it requires only polynomially many cells to
997: achieve a given fatness.  (The construction is also a polynomially effective
998: randomized algorithm.)  Given $N$, we find a strongly regular cellulation of
999: $S^3$ with $O(N^{12})$ cells and fatness at least $N$.  Note that there are
1000: also power-law upper bounds on fatness:  An $O(N^{1/3})$ upper bound for the
1001: fatness of a convex $4$-polytope with $N$ vertices follows from work by
1002: Edelsbrunner and Sharir \cite{ES:hyperplane}.  The K\H{o}vari-S\'os-Tur\'an
1003: theorem \cite{KST:problem} (see also \cite[p.~1239]{Bollobas:extremal} and
1004: \cite[Thm.~9.6,p.~121]{PA:combin}) implies an $O(N^{2/3})$ upper bound on the
1005: fatness of strongly regular cellulations of $S^3$, since the vertex-facet
1006: (atom-coatom) incidence graph has no $K_{3,3}$-subgraph, and thus has at most
1007: $O\big(f_0(f_0+f_3)^{2/3}\big)$ edges. Our construction provides an
1008: $\Omega(N^{1/12})$ lower bound.
1009: 
1010: The inefficient construction is a simpler version which we describe first.  For
1011: every $g>0$, $S^3$ can be realized as
1012: \[
1013: H_1 \cup (S_g \times I) \cup H_2,
1014: \]
1015: a thickened surface of genus $g$ capped on both ends with handlebodies. (This
1016: is obtained from a neighborhood of the standard [unknotted] smooth embedding
1017: of $S_g$ into $S^3$.) If for some $g$ we can find a fat cellulation of $S_g$,
1018: we can realize $S^3$ as a ``fat sausage with lean ends,'' as shown in
1019: \fig{f:sausage}.  We cross the fat cellulation of $S_g$ with an interval
1020: divided into $N$ segments to produce a fat cellulation of $S_g \times I$.  Then
1021: we fix arbitrary strongly regular cellulations of the handlebodies $H_1$ and
1022: $H_2$.  If we make the sausage long enough, \ie, if we take $N \to \infty$, the
1023: fatness of the sausage converges to the fatness of its middle regardless of the
1024: structure of its ends.
1025: 
1026: \begin{fullfigure}{f:sausage}{$S^3$ as a fat sausage with lean ends}
1027: \begin{large} \psset{unit=.2in} \pspicture(-8,-2.5)(8,3)
1028: \multips(-6,0)(4,0){4}{
1029:     \psbezier(.7,-1)(.7,-1.5)(.35,-2)(0,-2)
1030:     \psbezier(.7,-1)(.7,-.5)(.35,-.5)(.35,0)
1031:     \psbezier(.35,0)(.35,.5)(.7,.5)(.7,1)
1032:     \psbezier(.7,1)(.7,1.5)(.35,2)(0,2)
1033:     \psbezier(-.7,-1)(-.7,-1.5)(-.35,-2)(0,-2)
1034:     \psbezier(-.7,-1)(-.7,-.5)(-.35,-.5)(-.35,0)
1035:     \psbezier(-.35,0)(-.35,.5)(-.7,.5)(-.7,1)
1036:     \psbezier(-.7,1)(-.7,1.5)(-.35,2)(0,2)
1037:     \psbezier(0,.65)(.25,.9)(.25,1.15)(0,1.4)
1038:     \psbezier(0,.65)(-.25,.9)(-.25,1.15)(0,1.4)
1039:     \psbezier(0,-.65)(.25,-.9)(.25,-1.15)(0,-1.4)
1040:     \psbezier(0,-.65)(-.25,-.9)(-.25,-1.15)(0,-1.4)}
1041: \psline(-6,2)(6,2)\psline(-6,-2)(6,-2)
1042: \psarc(6,0){2}{270}{90}\psarc(-6,0){2}{90}{270}
1043: \psline[linestyle=dashed](-6,1.4)(6,1.4)
1044: \psline[linestyle=dashed](-6,.65)(6,.65)
1045: \psline[linestyle=dashed](-6,-1.4)(6,-1.4)
1046: \psline[linestyle=dashed](-6,-.65)(6,-.65)
1047: \rput(-7.25,0){$H_1$}\rput(0,0){$S_g \times I$}\rput(7.25,0){$H_2$}
1048: \endpspicture \end{large} \eatline \end{fullfigure}
1049: 
1050: It remains only to show that there are strongly regular fat cellulations of
1051: surfaces.  The surface $S_g$ has perfect cellulations with $f$-vector
1052: $(1,2g,1)$.  Such a cellulation is obtained by gluing pairs of sides of a
1053: $4g$-gon in such a way that all vertices are identified. It has fatness $g$,
1054: and it exists for arbitrarily large $g$, but it is far from regular. However,
1055: its lift to the universal cover $\tilde{S}_g$ is strongly regular, since any
1056: such cellulation can be represented by a tiling of the hyperbolic plane by
1057: \emph{convex} polygons.  (Indeed, if we take the regular $4g$-gon with angles
1058: of $\pi/(2g)$, which is certainly convex, then its edges and angles are
1059: compatible with any perfect cellulation.)  Moreover, Mal'cev's theorem
1060: \cite{Malcev:faithful}, states that finitely generated matrix groups are
1061: residually finite; this implies that every closed hyperbolic manifold admits
1062: intermediate finite covers with arbitrarily large injectivity radius.  (See
1063: \cite[\S4]{Kuperberg:saturated} for a detailed exposition.) In particular $S_g$
1064: admits an intermediate cover $\hS_g$ whose injectivity radius exceeds the
1065: diameter of a $2$-cell.  The cellulation of $\hS_g$ is then strongly
1066: regular.  Its genus is much larger than $g$, but its fatness is still $g$.
1067: 
1068: The efficient construction is the same: It only requires careful choices for
1069: the finite cover $\hS_g$ and for the handlebodies $H_1$ and $H_2$. Among
1070: the perfect cellulations of $S_g$, a convenient one for us is a $4g$-gon with
1071: opposite edges identified.  We describe the fundamental group $\pi_1(S_g)$
1072: using this cellulation.  As shown in \fig{f:sg}, we number the edges
1073: $x_0,\dots,x_{4g-1}$ consecutively, so that
1074: \eq{e:ident}{x_i^{} = x_{2g+i}^{-1}}
1075: and
1076: \eq{e:cell}{x_0x_1\cdots x_{4g-1} = 1.}
1077: (We interpret the indices as elements of $\Z/(4g)$.)
1078: Equation~\eqref{e:ident} expresses the identifications, while
1079: equation~\eqref{e:cell} expresses the boundary of the $2$-cell.
1080: 
1081: \begin{fullfigure}{f:sg}{Labelling edges of $S_g$}
1082: \pspicture(-4.5,-4.5)(4,4.5) \sgpie
1083: \uput[90](3.435;90){$x_0$} \uput[120](3.435;120){$x_1$}
1084: \uput[150](3.435;150){$x_2$} \uput[240](3.435;240){$x_{2g-1}$}
1085: \uput[270](3.435;270){$x_{2g}$} \uput[60](3.435;60){$x_{4g-1}$}
1086: \eatline \endpspicture \end{fullfigure}
1087: 
1088: We construct $\hS_g$ as a tower of two abelian finite covers, which
1089: together form an irregular cover of $S_g$. (No abelian cover of $S_g$ is
1090: strongly regular.) The surface $S_g$ satisfies the usual requirements of
1091: covering-space
1092: theory (see Fulton \cite[\S\ 13b, 14a]{Fulton:gtm}): It is a connected,
1093: locally path-connected, and locally simply connected space. For any finite
1094: group $A$, the \emph{$A$-coverings} of $S_g$ are covering spaces of the form
1095: $Y$ with $Y/A=S_g$, where $A$ acts properly discontinuously on $Y$. These
1096: coverings, up to isomorphism, are classified by the set of group homomorphisms
1097: $\Hom(\pi_1(S_g,x),A)$ \cite[Thm.~14.a]{Fulton:gtm}. Furthermore, if $A$ is
1098: abelian, then every such homomorphism maps the commutators in $\pi_1(S_g,x)$ to
1099: zero, so the $A$-coverings are classified by $\Hom(H_1(S_g,\Z),A)$. In other
1100: words, if $A$ is any abelian group with $n$ elements, then every homomorphism
1101: $\sigma: H_1(S_g)\rightarrow A$ defines an $n$-fold abelian covering of $S_g$.
1102: (If the homomorphism is not surjective, then the covering space is not
1103: connected \cite[p.~193]{Fulton:gtm}. In this case we use a connected component
1104: of the covering space, which has the same fatness but smaller genus.)
1105: 
1106: Now assume that $q = 4g+1$ is a prime power and let $\alpha$ generate the
1107: cyclic group $\F_q^*$, where $\F_q$ is the field with $q$ elements.  We can
1108: fulfill the assumption by changing $g$ by a bounded factor.  (Most simply we
1109: can let $q = 5^k$.  Or we can let $q$ be prime, so that $\F_q=\Z/q$, by a form
1110: of Bertrand's postulate for primes in congruence classes.  This result dates to
1111: the 19th century; see Erd\H{o}s \cite{Erdos:primzahlen} for an elementary
1112: proof.) Define a homomorphism $\sigma:H_1(S_g) \to \F_q$ by $\sigma([x_i]) =
1113: \alpha^i$, where $[x_i]$ is the 1-cycle (or homology class) represented by the
1114: loop $x_i$.  Since $\alpha^{2g} = -1$, the definition of $\sigma$ is consistent
1115: with equation~\eqref{e:ident}. Consistency with \eqref{e:cell} is then
1116: automatic. Let $S_g'$ be the finite cover corresponding to $\sigma$.
1117: 
1118: To prepare for the subsequent analysis of $\hS_g$, we give an explicit
1119: combinatorial description of $S'_g$.  Let $F^0$ be a $2$-cell of $S'_g$
1120: and label its vertices
1121: $$v_0^0,v_1^0,\ldots,v_{4g-1}^0$$
1122: in cyclic order.  See \fig{f:sg1}.
1123: 
1124: \begin{fullfigure}{f:sg1}{Labelling the vertices of $F^0$}
1125: \pspicture(-5,-4.75)(5,4.75) \sgpie \rput(0,0){\large $F^0$}
1126: \uput[90](3.435;90){$+1$} \uput[120](3.435;120){$+\alpha$}
1127: \uput[150](3.435;150){$+\alpha^2$}
1128: \uput{10pt}[270](3.435;270){$+\alpha^{2g} = -1$}
1129: \uput[300](3.435;300){$+\alpha^{2g} = -\alpha$}
1130: \uput[45](4;45){$v^0_{4g-1}$} \uput[75](4;75){$v^0_0$}
1131: \uput[105](4;105){$v^0_1$}
1132: \eatline \endpspicture \end{fullfigure}
1133: 
1134: For each $s \in \F_q$, let $F^s$ and $v_k^s$ be the images of $F^0$ and $v_k^0$
1135: under the action of $s$. Since $S_g$ has only one vertex, the vertices of
1136: $S_g'$ may be identified with $\F_q$. Thus, if we identify $v_0^0$ with
1137: $0\in\F_q$, then the action of $\F_q$ will identify $v_0^s$ with $s\in\F_q$.
1138: The structure of $\sigma$ further implies that
1139: \[
1140: v_k^s\ =\ s+1+\alpha+\alpha^2+\ldots+\alpha^{k-1}\ =\
1141:     s+\frac{\alpha^k-1}{\alpha-1}
1142: \]
1143: for all $k\in\Z/(4g)$ and $s\in\F_q$.  See \fig{f:sg2}.
1144: 
1145: \begin{fullfigure}{f:sg2}{Labelling the vertices of $F^s$}
1146: \pspicture(-7,-4.75)(5.5,4.75) \sgpie \rput(0,0){\large $F^s$}
1147: \uput[75](4;75){$v^s_0 = s$}
1148: \uput[105](4;105){$v^s_1 = s+1$}
1149: \uput[120](4;135){$v^s_2 = s+1+\alpha$}
1150: \uput[135](4;165){$v^s_3 = s+\tfrac{\alpha^3-1}{\alpha-1}$}
1151: \eatline \endpspicture \end{fullfigure}
1152: 
1153: Using this explicit description, it is routine to verify the following
1154: (remarkable) properties of the surface $S'_g$.
1155: 
1156: \begin{lemma}
1157: The cellulation of the abelian cover $S'_g$ is regular and has $f$-vector
1158: \[
1159: f(S'_g) = (q,2gq,q)=(q,\binom{q}{2},q).
1160: \]
1161: Every facet has $q-1=4g$ vertices, while every vertex has degree $q-1=4g$. The
1162: graph (or $1$-skeleton) of $S_g'$ is the complete graph on $q+1$ vertices. The
1163: dual graph is also complete; any two facets share exactly one edge (as well as
1164: $q-4$ other vertices).
1165: \label{l:regular} \end{lemma}
1166: 
1167: \begin{proof}
1168: In view of the combinatorial description above (\fig{f:sg2}), all these
1169: facts follow from simple computations in the field $\F_q$:
1170: \begin{itemize}\setlength{\itemsep}{0pt}
1171: \item $S_g'$ is regular --- for each $s\in\F_q$, the vertex labels
1172: $s+\frac{\alpha^k-1}{\alpha-1}$ ($0\le k<4g$) are distinct.
1173: \item The $1$-skeleton of $S_g'$ is complete --- for $v,v'\in\F_q$, $v\neq v'$
1174: there is a unique $s\in\F_q$ and $k\in\Z/(4g)$ with
1175: \[
1176: v=s+\frac{\alpha^k-1}{\alpha-1}\qquad
1177:     v'=s+\frac{\alpha^{k+1}-1}{\alpha-1}.
1178: \]
1179: \item The dual graph of $S_g'$ is complete --- for $s,s'\in\F_q$, $s\neq s'$,
1180: there are unique $k,\ell\in\Z/(4g)$ such that
1181: \begin{align*}
1182: s+\frac{\alpha^k-1}{\alpha-1}&=s'+\frac{\alpha^{\ell+1}-1}{\alpha-1} \\
1183: s+\frac{\alpha^{k+1}-1}{\alpha-1}&=s'+\frac{\alpha^\ell-1}{\alpha-1}.
1184: \end{align*}
1185: \end{itemize}
1186: \end{proof}
1187: 
1188: \begin{theorem}
1189: Let $n \ge 128 g^4$, and let
1190: \[
1191: \rho:H_1(S_g')\ \to\ \Z/n
1192: \]
1193: be a randomly chosen homomorphism, and let $\hS_g$ be the finite cover of
1194: $S_g'$ corresponding to $\rho$. Then with probability more than $\frac12$, the
1195: cellulation of\/ $\hS_g$ is strongly regular.
1196: \label{th:prob}
1197: \end{theorem}
1198: 
1199: In order to prove \thm{th:prob}, we need to more explicitly describe the
1200: condition of strong regularity as it applies to $\hS_g$. Let $X$ be a regular
1201: cell complex and suppose that its universal cover $\tilde{X}$ is strongly
1202: regular.  Recall that the \emph{star} $\st(v)$ of a vertex $v$ in $X$ is the
1203: subcomplex generated by the cells that contain $v$. The complex $X$ is strongly
1204: regular if and only if the star of each vertex is.  Suppose that $\tilde{v} \in
1205: \tilde{X}$ projects to $v\in X$. Then the star $\st(\tilde{v})$, which is
1206: strongly regular, projects to the star $\st(v)$.  The latter is strongly
1207: regular if and only if the projection is injective.  In other words, $X$ is
1208: strongly regular if and only if the stars of $\tilde{X}$ embed in $X$. If $X$
1209: is not strongly regular, then there must be a path $\tilde{\ell}$ connecting
1210: distinct vertices of $\st(\tilde{v})$ which projects to a loop $\ell$ in
1211: $\st(v)$.  We say that such a loop \emph{obstructs strong regularity}.  We can
1212: assume that $\tilde{\ell}$ is a pair of segments properly embedded in distinct
1213: cells in $\st(\tilde v)$, with only the end-points of the segments on the
1214: boundary of the cells, which implies that $\ell$ is embedded if $X$ is regular.
1215: \fig{f:obstruct} gives an example of such a loop $\ell$ in a regular
1216: cellulation of a torus.
1217: 
1218: \begin{fullfigure}{f:obstruct}{A loop $\ell$ that obstructs strong regularity
1219:     in the torus $S_1'$}
1220: \pspicture(0,0)(6,4)
1221: \pspolygon[linestyle=none,fillstyle=solid,fillcolor=gray90]
1222:     (0,0)(0,2)(2,2)(2,4)(6,4)(6,0)
1223: \psline(0,0)(6,0) \psline(0,2)(6,2) \psline(2,4)(6,4) \psline(0,0)(0,2)
1224: \psline(2,0)(2,4) \psline(4,0)(4,4) \psline(6,0)(6,4)
1225: \multips(0,0)(2,0){4}{\qdisk(0,0){.1}} \multips(0,2)(2,0){4}{\qdisk(0,0){.1}}
1226: \multips(2,4)(2,0){3}{\qdisk(0,0){.1}}
1227: \rput(1,1){\large $F^1$} \rput(3,1){\large $F^0$} \rput(5,1){\large $F^4$}
1228: \rput(3,3){\large $F^3$} \rput(5,3){\large $F^2$}
1229: \uput[225](0,0){$4$} \uput[270](2,0){$3$} \uput[270](4,0){$2$}
1230: \uput[315](6,0){$1$}
1231: \uput[135](0,2){$2$} \uput[135](2,2){$1$} \uput[135](4,2){$0$}
1232: \uput[45](6,2){$4$}
1233: \uput[135](2,4){$4$} \uput[90](4,4){$3$} \uput[45](6,4){$2$}
1234: \pscurve[linestyle=dashed](2,0)(2.8,.4)(3.6,.8)(3.4,1.6)(4,2)(4.4,3)
1235:     (4.6,3.6)(4,4)
1236: \rput(3.7,.4){\large $\ell$}
1237: \endpspicture
1238: \end{fullfigure}
1239: 
1240: In our case, the surfaces $S_g'$ are regular, but they have many obstructing
1241: loops.  \thm{th:prob} asserts that, with non-zero probability, all such loops
1242: lengthen when lifted to $\hS_g$.
1243: 
1244: \begin{lemma} No loops in $S_g'$ that obstruct strong regularity are
1245: null-homologous.  Furthermore, all obstructing loops represent indivisible
1246: elements in $H_1(S_g')$.
1247: \label{l:notnull} \end{lemma}
1248: 
1249: \begin{proof} In brief, they are indivisible because they are embedded,
1250: and they are too short to be null-homologous.
1251: 
1252: By \lem{l:regular}, $S_g'$ is regular.  By the discussion after the statement
1253: of \thm{th:prob}, each obstructing loop $\ell$ is embedded. If $\ell$ separates
1254: $S_g'$, then it is null-homologous. If $\ell$ does not separate $S_g'$, then it
1255: is indivisible in homology.  (To show this, we can appeal to the classification
1256: of surfaces by cutting $S_g'$ along $\ell$.  The classification implies that
1257: all non-separating positions for $\ell$ are equivalent up to homeomorphism of
1258: $S_g'$.  It is easy to find a standard position for $\ell$ in which it is
1259: indivisible in homology.) Thus it remains to show that no obstructing loop is
1260: null-homologous.
1261: 
1262: First, we claim that any obstructing loop $\ell$ can be supported on fewer than
1263: $4g$ edges of the $1$-skeleton of $S_g'$.  We homotop the two segments of
1264: $\ell$ to the boundaries of the 2-cell containing them, giving them each at
1265: most $2g$ edges.  Thus $\ell$ is represented by a sequence of at most $4g$
1266: edges in $S_g'$. The case of exactly $4g$ edges does not occur, since the
1267: endpoints of the loop coincide, and no two vertices $v\neq v'$ of $S_g'$ are
1268: opposite vertices in two different facets $F^s$.  As in \lem{l:regular}, this
1269: follows from the fact that for $v,v'\in\F_g$, $v\neq v'$, there are unique
1270: $s\in\F_q$ and $k\in\Z/(4g)$ with
1271: \[
1272: v = s+\frac{\alpha^k-1}{\alpha-1}\qquad
1273: v' = s+\frac{\alpha^{k+2g}-1}{\alpha-1}.
1274: \]
1275: 
1276: Second,
1277: we claim that any null-homologous loop in the $1$-skeleton
1278: of $S_g'$ contains at least $4g$ edges.  In other we if $f$ is a $2$-chain on
1279: $S_g'$ and $\d f \ne 0$, then $|\d f| \ge 4g$. Since $S_g'$ is orientable, we
1280: can regard $f$ as a function on its $2$-cells.  Since $f$ is non-constant, it
1281: attains some value $t$ on $k$ $2$-cells with $0 < k < q$.  Since any two
1282: $2$-cells share an edge by \lem{l:regular}, these $2$-cells share
1283: \[
1284: k(q-k) \ge 4g
1285: \]
1286: edges with the complementary set of $2$-cells, of which there are $q-k$.  Since
1287: $\d f$ is non-zero on these edges, $|\d f| \ge 4g$, as desired.
1288: \end{proof}
1289: 
1290: \begin{proof}[Proof of \thm{th:prob}] In brief, $S_g'$ has fewer than $64g^4$
1291: obstructing loops $\ell$.  For each one,
1292: \[
1293: P\bigl[\rho([\ell]) = 0\bigr] = \frac1n.
1294: \]
1295: The expected number of obstructing loops that lift from $S_g'$ to $\hS_g$
1296: without lengthening is less than $64g^4/n \le \frac12$.  Thus there is a good
1297: chance that all obstructing loops lengthen.
1298: 
1299: The homology group $H_1(S'_g)$ is a finitely generated free abelian group: It
1300: is isomorphic to $\Z^d$, for $d=1+q(g-1)$. Thus it admits $n^d$ homomorphisms
1301: $\rho$ to $\Z/n$.  Since this is a finite number, choosing one uniformly at
1302: random is well-defined.  If $c$ is any indivisible vector in $\Z^d$, then it is
1303: contained in a basis, and thus $\rho(c)$ is equidistributed. In particular, if
1304: $\ell$ is an obstructing loop, then $[\ell]$ is indivisible by \lem{l:notnull},
1305: so $\rho([\ell])$ is equidistributed in $\Z/n$.
1306: 
1307: It remains only to bound the number of obstructing loops in $S_g'$. A star
1308: $\st(v)$ in $\tilde{S}_g$ has $4g(4g-2)$ points other than $v$ itself.
1309: Without loss of generality, $v$ projects to $0$ in $S_g'$.  In this case the
1310: other vertices are equidistributed among the $4g$ non-zero elements of
1311: $\F_q$.  Therefore $\st(v)$ has
1312: \[
1313: 4g\binom{4g-2}{2}\ =\ \frac{4g(4g-2)(4g-3)}{2}
1314: \]
1315: pairs of arcs connecting $v$ to two vertices that are the same in $S_g'$.
1316: These pairs represent all two-segment obstructing loops that pass through $0$
1317: and a nonzero vertex $v'$ (and some of these loops are homotopic).  If we count
1318: such pairs of arcs for any pair of distinct vertices $v,v'$
1319: of $S_g'$, then we find that the total number is not more than
1320: \[
1321: \binom{4g+1}{2}\binom{4g-2}{2}\ =\
1322: \frac{(4g+1)4g(4g-2)(4g-3)}{4} < 64g^4,
1323: \]
1324: as desired.
1325: \end{proof}
1326: 
1327: \begin{question} For each $g>1$, what is the maximum fatness of a strongly
1328: regular cellulation of a surface of genus $g$? Equivalently, how many edges are
1329: needed for a strongly regular cellulation of a surface of genus $g$?
1330: \end{question}
1331: 
1332: \begin{remark}
1333: One interesting alternative to the construction of $S_g'$ is to assume instead
1334: that $q = 4g-1$ is a prime power, and to let $\alpha$ be an element of order
1335: $4g$ in $\F_{q^2}$.  The resulting $q^2$-fold cover $S_g''$ is almost strongly
1336: regular: the only obstructing loops are those that are null homologous in
1337: $S_g$.  Another interesting surface is the modular curve $X(2p)$, where $p$ is
1338: a prime \cite[\S13]{Silverman:gtm}. The inclusion $\Gamma(2p) \subset
1339: \Gamma(2)$ of modular groups induces a projection from $X(2p)$ to the modular
1340: curve $X(2)$, which is a sphere with three cusp points. If we connect two of
1341: these points by an arc which avoids the third, it lifts to a cellulation of
1342: $X(2p)$ with $f$-vector $(p^2-1,\frac{p(p^2-1)}{2},\frac{p^2-1}{2})$. Like
1343: $S_g''$, it has a few obstructing loops. Unfortunately we do not know a way to
1344: use either $S_g''$ or $X(2p)$ to make fat surfaces of lower genus (or
1345: equivalently fewer cells) than $\hS_g$.
1346: \end{remark}
1347: 
1348: Since \thm{th:prob} provides us with efficient fat surfaces $\hS_g$, the
1349: construction of fat cellulations of $S^3$ only requires efficient cellulations
1350: of the handlebodies $H_1$ and $H_2$ and an efficient way to attach them to
1351: $\hS_g$. In our construction the handlebody cellulations are \emph{a priori}
1352: unrelated to the cellulation of $\hS_g$.  Rather they are transverse after
1353: attachment, and each point of intersection will become a new vertex.  Thus the
1354: question is to position the cellulations to minimize their intersection.
1355: 
1356: We describe the cellulations in three stages:  first, a
1357: dissection of $H_1$ and $H_2$ individually into $3$-cells; second, their
1358: relative position; and third, their position relative to the cellulation of
1359: $\hS_g$.
1360: 
1361: Let $\hg$ be the genus of $\hS_g$.  A handlebody $H$ of genus $\hg$ can
1362: be formed by identifying $\hg$ pairs of disks on the surface of a $3$-cell.
1363: The result is a dissection of $H$ into $\hg$ $2$-cells and one $3$-cell,
1364: although it is not a cell complex because there are no $1$-cells or $0$-cells.
1365: We can still ask whether such a dissection is regular or strongly regular; this
1366: one is neither.  However, if we replace each $2$-cell by 3 parallel $2$-cells,
1367: it becomes strongly regular.  An example of the resulting dissection $A$ is
1368: shown in \fig{f:tripled}.
1369: 
1370: \begin{fullfigure}{f:tripled}
1371:     {A strongly regular dissection $A$ of a handlebody of genus 2}
1372: \pspicture(-4,-2)(4,3)
1373: \psbezier(-2,1.4)(-3,1.4)(-4,0.7)(-4,0)
1374: \psbezier(-2,1.4)(-1,1.4)(-1,0.7)(0,0.7)
1375: \psbezier(0,0.7)(1,0.7)(1,1.4)(2,1.4)
1376: \psbezier(2,1.4)(3,1.4)(4,0.7)(4,0)
1377: \psbezier(-2,-1.4)(-3,-1.4)(-4,-0.7)(-4,0)
1378: \psbezier(-2,-1.4)(-1,-1.4)(-1,-0.7)(0,-0.7)
1379: \psbezier(0,-0.7)(1,-0.7)(1,-1.4)(2,-1.4)
1380: \psbezier(2,-1.4)(3,-1.4)(4,-0.7)(4,0)
1381: \psbezier(1.3,0)(1.8,0.5)(2.3,0.5)(2.8,0)
1382: \psbezier(1.3,0)(1.8,-0.5)(2.3,-0.5)(2.8,0)
1383: \psbezier(-1.3,0)(-1.8,0.5)(-2.3,0.5)(-2.8,0)
1384: \psbezier(-1.3,0)(-1.8,-0.5)(-2.3,-0.5)(-2.8,0)
1385: \psbezier(-2.025,.375)(-1.575,.375)(-1.575,1.4)(-2.025,1.4)
1386: \psbezier[linestyle=dashed](-2.025,.375)(-2.475,.375)(-2.475,1.4)(-2.025,1.4)
1387: \psbezier(-2.425,.3)(-1.975,.3)(-1.975,1.35)(-2.425,1.35)
1388: \psbezier[linestyle=dashed](-2.425,.3)(-2.875,.3)(-2.875,1.35)(-2.425,1.35)
1389: \psbezier(-1.625,.3)(-1.175,.3)(-1.175,1.35)(-1.625,1.35)
1390: \psbezier[linestyle=dashed](-1.625,.3)(-2.075,.3)(-2.075,1.35)(-1.625,1.35)
1391: \psbezier[linestyle=dashed](2.025,.375)(1.575,.375)(1.575,1.4)(2.025,1.4)
1392: \psbezier(2.025,.375)(2.475,.375)(2.475,1.4)(2.025,1.4)
1393: \psbezier[linestyle=dashed](2.425,.3)(1.975,.3)(1.975,1.35)(2.425,1.35)
1394: \psbezier(2.425,.3)(2.875,.3)(2.875,1.35)(2.425,1.35)
1395: \psbezier[linestyle=dashed](1.625,.3)(1.175,.3)(1.175,1.35)(1.625,1.35)
1396: \psbezier(1.625,.3)(2.075,.3)(2.075,1.35)(1.625,1.35)
1397: \endpspicture \end{fullfigure}
1398: 
1399: The surface $S_\hg$ (which for our choice of $\hg$ is isomorphic to
1400: $\hS_g$) has another standard perfect cellulation called a \emph{canonical
1401: schema} in the computer science literature \cite{VY:surfaces}.  Using the
1402: labelling in \fig{f:sg}, we identify $x_{4k}^{}$ with $x_{4k+2}^{-1}$ (the even
1403: loops), and $x_{4k+1}^{}$ with $x_{4k+3}^{-1}$ (the odd loops), for all $0 \le
1404: k \le \hg$. By rounding corners we can make each even loop intersect one odd
1405: loop once and eliminate all other intersections between loops. If we interpret
1406: this pattern of loops as the standard Heegaard diagram for $S^3$
1407: \cite{Stillwell:gtm}, then the even loops bound disks in $H_1$ and the odd
1408: loops bound disks in $H_2$.  We can then put in two copies $A_1$ and $A_2$ of
1409: the cell division $A$ so that its $2$-cells run parallel to these loops.
1410: 
1411: We would like to position $A_1$ and $A_2$ to minimize their intersection with
1412: the cellulation of $\hS_g$. (Note that $A_1$ and $A_2$ do not intersect each
1413: other since $\hS_g \times I$ lies in between.) To this end Vegter and Yap
1414: \cite{VY:surfaces} proved that any cellulation of $S_\hg$ with $n$ edges admits
1415: a position of the canonical schema cellulation with $O(n\hg)$ intersections.
1416: (Strictly speaking the theorem applies to triangulations, but any regular
1417: cellulation of a surface with $n$ edges can be refined to a triangulation with
1418: less than $3n$ edges.) In our case
1419: \[
1420: n = \Theta(\hg) = \Theta(g^6).
1421: \]
1422: Thus, by tripling the edges of the canonical schema in the Vegter-Yap
1423: construction, we can position $A_1$ and $A_2$ so that the lean ends in the
1424: sausage have $O(g^{12})$ vertices.  If we give the fat part of the sausage $N =
1425: g^{12}$ slices, the total $f$-vector of the cellulation of $S^3$ is then
1426: \[
1427: (\Theta(g^{12}),\Theta(g^{13}),\Theta(g^{13}),\Theta(g^{12})),
1428: \]
1429: and its fatness is $\Theta(g)$.  This completes the efficient construction
1430: with unbounded fatness.
1431: 
1432: % \bibliography{me,co,gt,books}
1433: 
1434: \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
1435: \begin{thebibliography}{10}
1436: 
1437: \bibitem{Barnette:inequalities}
1438: David Barnette, \emph{Inequalities for $f$-vectors of $4$-polytopes}, Israel J.
1439:   Math. \textbf{11} (1972), 284--291.
1440: 
1441: \bibitem{Bayer:extended}
1442: Margaret~M. Bayer, \emph{The extended $f$-vectors of $4$-polytopes}, J.
1443:   Combinatorial Theory Ser. A \textbf{44} (1987), no.~1, 141--151.
1444: 
1445: \bibitem{BJ:dehn}
1446: Margaret~M. Bayer and Louis~J. Billera, \emph{Generalized {Dehn-Sommerville}
1447:   relations for polytopes, spheres and {Eulerian} partially ordered sets},
1448:   Inventiones Math. \textbf{79} (1985), no.~1, 143--157.
1449: 
1450: \bibitem{BL:sufficiency}
1451: Louis~J. Billera and Carl~W. Lee, \emph{A proof of the sufficiency of
1452:   {M}c{M}ullen's conditions for $f$-vectors of simplicial convex polytopes}, J.
1453:   Combin. Theory Ser. A \textbf{31} (1981), no.~3, 237--255.
1454: 
1455: \bibitem{Bollobas:extremal}
1456: B\'ela Bollob\'as, \emph{Extremal graph theory}, Handbook of Combinatorics
1457:   (R.~Graham, M.~Gr\"otschel, and L.~Lov\'asz, eds.), North-Holland/Elsevier,
1458:   Amsterdam, 1995, pp.~1231--1292.
1459: 
1460: \bibitem{Braden:glued}
1461: Tom Braden, \emph{A glued hypersimplex}, 1997, Personal communication.
1462: 
1463: \bibitem{CS:splag}
1464: John~H. Conway and Neil J.~A. Sloane, \emph{Sphere packings, lattices and
1465:   groups}, 3rd ed., Grundlehren der mathematischen {Wissenschaften}, vol. 290,
1466:   Springer-Verlag, New York, 1993.
1467: 
1468: \bibitem{Coxeter:regpoly}
1469: Harold Scott~MacDonald Coxeter, \emph{Regular polytopes}, 2nd ed., Macmillan,
1470:   New York, 1963, Corrected reprint, Dover, New York 1973.
1471: 
1472: \bibitem{ES:hyperplane}
1473: Herbert Edelsbrunner and Micha Sharir, \emph{A hyperplane incidence problem
1474:   with applications to counting distances}, Applied geometry and discrete
1475:   mathematics: {The Victor Klee Festschrift} (P.~Gritzman and B.~Sturmfels,
1476:   eds.), DIMACS Series in Discrete Mathematics and Theoretical Computer
1477:   Science, vol.~4, Amer. Math. Soc., Providence, RI, 1991, pp.~253--263.
1478: 
1479: \bibitem{Erdos:primzahlen}
1480: Paul Erd\H{o}s, \emph{{\"Uber} die {Primzahlen} gewisser arithmetischer
1481:   {Reihen}}, Math. Z. \textbf{39} (1935), 473--491.
1482: 
1483: \bibitem{Kuperberg:saturated}
1484: G\'abor {Fejes T\'oth}, Greg Kuperberg, and W{\l}odzimierz Kuperberg,
1485:   \emph{Highly saturated packings and reduced coverings}, Monatsh. Math.
1486:   \textbf{125} (1998), no.~2, 127--145.
1487: 
1488: \bibitem{Fulton:gtm}
1489: William Fulton, \emph{Algebraic topology. {A} first course}, Graduate Texts in
1490:   Mathematics, vol. 153, Springer-Verlag, New York, 1995.
1491: 
1492: \bibitem{Grunbaum:convex}
1493: Branko Gr\"unbaum, \emph{Convex polytopes}, Interscience, London, 1967.
1494: 
1495: \bibitem{HZ:census}
1496: Andrea H\"oppner and G\"unter~M. Ziegler, \emph{A census of flag-vectors of
1497:   $4$-polytopes}, Polytopes --- combinatorics and computation (G.~Kalai and
1498:   G.M. Ziegler, eds.), DMV Seminars, vol.~29, Birkh\"auser-Verlag, Basel, 2000,
1499:   pp.~105--110.
1500: 
1501: \bibitem{Iverson:hyperbolic}
1502: Birger Iversen, \emph{Hyperbolic geometry}, London Math. Soc. Student Texts,
1503:   vol.~25, Cambridge University Press, Cambridge, 1992.
1504: 
1505: \bibitem{JZ:cubical}
1506: Michael Joswig and G\"unter~M. Ziegler, \emph{Neighborly cubical polytopes},
1507:   Discrete Comput. Geom. \textbf{24} (2000), no.~2-3, 325--344,
1508:   \mbox{arXiv:math.CO/9812033}.
1509: 
1510: \bibitem{KST:problem}
1511: Tam\'as K\H{o}v\'ari, Vera~T. S\'os, and P\'al Tur\'an, \emph{On a problem of
1512:   {K. Zarankiewicz}}, Colloq. Math. \textbf{3} (1954), 50--57.
1513: 
1514: \bibitem{Kuperberg:kissing}
1515: Greg Kuperberg and Oded Schramm, \emph{Average kissing numbers for
1516:   non-congruent sphere packings}, Math. Res. Lett. \textbf{1} (1994), no.~3,
1517:   339--344, \mbox{arXiv:math.MG/9405218}.
1518: 
1519: \bibitem{Malcev:faithful}
1520: A.~I. Mal'cev, \emph{On the faithful representation of infinite groups by
1521:   matrices}, Amer. Math. Soc. Transl. (2) \textbf{45} (1965), 1--18.
1522: 
1523: \bibitem{Malkevitch:tiling}
1524: Joseph Malkevitch, \emph{Tiling convex polygons with equilateral triangles and
1525:   squares}, Discrete Geometry and Convexity (J.~E. Goodman, E.~Lutwak,
1526:   J.~Malkevitch, and R.~Pollack, eds.), Ann. New York Acad. Sci., vol. 440, New
1527:   York Academy of Sciences, 1985, pp.~299--303.
1528: 
1529: \bibitem{McMullen:simplicial}
1530: Peter McMullen, \emph{The numbers of faces of simplicial polytopes}, Israel J.
1531:   Math. \textbf{9} (1971), 559--570.
1532: 
1533: \bibitem{Munkres:algtop}
1534: James~R. Munkres, \emph{Elements of algebraic topology}, Addison-Wesley, Menlo
1535:   Park, CA, 1984.
1536: 
1537: \bibitem{PA:combin}
1538: J\'anos Pach and Pankaj~K. Agarwal, \emph{Combinatorial geometry}, J. Wiley and
1539:   Sons, New York, 1995.
1540: 
1541: \bibitem{Schulte:steinitz}
1542: Egon Schulte, \emph{Analogues of {Steinitz's} theorem about
1543:   non-in\-scrib\-a\-ble polytopes}, Intuitive geometry (Si\'ofok 1985)
1544:   (Amsterdam), Colloquia Soc. J\'anos Bolyai, vol.~48, North Holland, 1987,
1545:   pp.~503--516.
1546: 
1547: \bibitem{Silverman:gtm}
1548: Joseph~H. Silverman, \emph{The arithmetic of elliptic curves}, Graduate Texts
1549:   in Mathematics, vol. 106, Springer-Verlag, New York, 1986.
1550: 
1551: \bibitem{Stanley:simplicial}
1552: Richard~P. Stanley, \emph{The number of faces of a simplicial convex polytope},
1553:   Adv. in Math. \textbf{35} (1980), no.~3, 236--238.
1554: 
1555: \bibitem{Stanley:eulerian}
1556: \bysame, \emph{A survey of {Eulerian} posets}, Polytopes: abstract, convex and
1557:   computational (Scarborough, ON, 1993), NATO Adv. Sci. Inst. Ser. C Math.
1558:   Phys. Sci., vol. 440, Kluwer Acad. Publ., Dordrecht, 1994, pp.~301--333.
1559: 
1560: \bibitem{Steinitz:eulerschen}
1561: Ernst Steinitz, \emph{{\"Uber} die {Eulerschen Polyederrelationen}}, Archiv
1562:   f\"ur Mathematik und Physik \textbf{11} (1906), 86--88.
1563: 
1564: \bibitem{Stillwell:gtm}
1565: John Stillwell, \emph{Classical topology and combinatorial group theory}, 2nd
1566:   ed., Graduate Texts in Mathematics, no.~72, Springer-Verlag, 1993.
1567: 
1568: \bibitem{Thurston:geometry}
1569: William~P. Thurston, \emph{Three-dimensional geometry and topology}, vol.~I,
1570:   Princeton University Press, Princeton, NJ, 1997.
1571: 
1572: \bibitem{VY:surfaces}
1573: Gert Vegter and Chee~K. Yap, \emph{Computational complexity of combinatorial
1574:   surfaces}, Proc. 6th ACM Symp. Comp. Geom., ACM Press, 1990, pp.~102--111.
1575: 
1576: \bibitem{Vinberg:reflection}
1577: Ernest~B. Vinberg, \emph{Hyperbolic reflection groups}, Russian Math. Surveys
1578:   \textbf{40} (1985), 31--75.
1579: 
1580: \bibitem{Waite:combining}
1581: William Waite, \emph{Combining squares and triangles}, Cubism for Fun
1582:   \textbf{53} (2000), 18--22.
1583: 
1584: \bibitem{Walkup:lower}
1585: David~W. Walkup, \emph{The lower bound conjecture for $3$- and $4$-manifolds},
1586:   Acta Math. \textbf{125} (1970), 75--107.
1587: 
1588: \bibitem{Ziegler:gtm}
1589: G\"unter~M. Ziegler, \emph{Lectures on polytopes}, Graduate Texts in
1590:   Mathematics, vol. 152, Springer-Verlag, New York, 1995, Revised edition,
1591:   1998.
1592: 
1593: \end{thebibliography}
1594: 
1595: 
1596: \end{document}
1597: