math0010302/p2.tex
1: % revision as suggested by referee of DCG, March 2, 2005. 
2: %
3: % Part II, 9 jun 04 (rlg corrections)
4: % 8 jun 04 
5: % 5 jun 04 ( cyan Thm 5.2 replace)
6: % new file from scratch 2 may 04 [April 15 last version]
7: %
8: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
9: % 11 Figure files 
10: %
11: %  0.0.1.1.ps 1.2.2.3.ps 6.11.14.15.ps
12: %  super.ps   
13: %  all6.ps all8.ps all9.ps 
14: %  0mod1.ps
15: %  1mo2.ps 2mod4.ps 0.mod4.ps
16: %   fig1.1a, 1-std.ps, 6-std.p
17: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  New results added since 2001 version:
18: %
19: %    Theorem 3.1, Theorem 3.2 (geometric ``packing''
20: %    Theorem 4.1 (restated, not new otherwise)
21: %    Theorem 5.1 (restated, not new otherwise)
22: %    Theoem 6.2, Theorem 6.3
23: %    Theorem 8.1 (CLM)
24: %    
25: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
26: \documentclass[11pt]{article}
27: %\usepackage{showlabels}
28: 
29: \usepackage{epsfig,amsmath,latexsym}
30: \usepackage{amsfonts,amssymb}
31: %\usepackage{diagrams}
32: 
33: 
34: \setlength{\textwidth}{6.2in}
35: \setlength{\textheight}{8.5in}
36: \setlength{\oddsidemargin}{.2in}
37: \setlength{\topmargin}{-0.25in}
38: \setlength{\headheight}{0in}
39: 
40: \newtheorem{theorem}{Theorem}[section]
41: \newtheorem{cor}[theorem]{Corollary}
42: \newtheorem{defi}{Definition}[section]
43: \newtheorem{lemma}[theorem]{Lemma}
44: \newtheorem{prop}[theorem]{Proposition}
45: \newtheorem{conj}{Conjecture}[section]
46: 
47: 
48: \def\binom#1#2{{#1}\choose{#2}}
49: 
50: \def\slfrac#1#2{\hbox{\kern.1em %
51:  \raise.5ex\hbox{\the\scriptfont0 #1}\kern-.11em %
52:  /\kern-.15em\lower.25ex\hbox{\the\scriptfont0 #2}}}
53: 
54: 
55: \newcommand{\ep}{\epsilon}
56: \newcommand{\pf}{\noindent{\bf Proof.~}}
57: \newcommand{\beq}{\begin{eqnarray}}
58: \newcommand{\eeq}{\end{eqnarray}}
59: \newcommand{\beql}[1]{\begin{eqnarray}\label{#1}}
60: \newcommand{\beqs}{\begin{eqnarray*}}
61: \newcommand{\eeqs}{\end{eqnarray*}}
62: \newcommand{\eqn}[1]{(\ref{#1})}
63: 
64: \newcommand{\bdelta}{\mbox{\boldmath $\delta$}}
65: \newcommand{\bj}{\mathbf j}
66: \newcommand{\bone}{\mathbf 1}
67: \newcommand{\lf}{\lfloor}
68: \newcommand{\rf}{\rfloor}
69: \newcommand{\lc}{\lceil}
70: \newcommand{\rc}{\rceil}
71: \newcommand{\cc}{{\mathbb C}}
72: \newcommand{\rr}{{\mathbb R}}
73: \newcommand{\zz}{{\mathbb Z}}
74: \newcommand{\qq}{{\mathbb Q}}
75: \newcommand{\nn}{{\mathbb N}}
76: \newcommand{\dd}{{\mathbb D}}
77: 
78: \newcommand{\CC}{{\mathbb C}}
79: \newcommand{\DD}{{\mathbb D}}
80: \newcommand{\RR}{{\mathbb R}}
81: \newcommand{\ZZ}{{\mathbb Z}}
82: \newcommand{\QQ}{{\mathbb Q}}
83: 
84: \newcommand{\fc}{{\mathfrak c}}
85: \newcommand{\fd}{{\mathfrak d}}
86: \newcommand{\fg}{{\mathfrak g}}
87: \newcommand{\fh}{{\mathfrak h}}
88: \newcommand{\fj}{{\mathfrak j}}
89: \newcommand{\fk}{{\mathfrak k}}
90: \newcommand{\fp}{{\mathfrak p}}
91: \newcommand{\fq}{{\mathfrak q}}
92: \newcommand{\fr}{{\mathfrak r}}
93: \newcommand{\fs}{{\mathfrak s}}
94: \newcommand{\ft}{{\mathfrak t}}
95: \newcommand{\fx}{{\mathfrak x}}
96: \newcommand{\fy}{{\mathfrak y}}
97: 
98: \newcommand{\ba}{{\mathbf a}}
99: \newcommand{\bc}{{\mathbf c}}
100: \newcommand{\br}{{\mathbf r}}
101: \newcommand{\be}{{\mathbf e}}
102: \newcommand{\bg}{{\mathbf g}}
103: \newcommand{\bh}{{\mathbf h}}
104: \newcommand{\bo}{{\mathbf 1}}
105: \newcommand{\bs}{{\mathbf s}}
106: \newcommand{\bv}{{\mathbf v}}
107: \newcommand{\bx}{{\mathbf x}}
108: \newcommand{\bz}{{\mathbf z}}
109: 
110: \newcommand{\bA}{{\mathbf A}}
111: \newcommand{\bB}{{\mathbf B}}
112: \newcommand{\bC}{{\mathbf C}}
113: \newcommand{\bD}{{\mathbf D}}
114: \newcommand{\bG}{{\mathbf G}}
115: \newcommand{\bI}{{\mathbf I}}
116: \newcommand{\bJ}{{\mathbf J}}
117: \newcommand{\bM}{{\mathbf M}}
118: \newcommand{\bP}{{\mathbf P}}
119: \newcommand{\bQ}{{\mathbf Q}}
120: \newcommand{\bS}{{\mathbf S}}
121: \newcommand{\bT}{{\mathbf T}}
122: \newcommand{\bU}{{\mathbf U}}
123: \newcommand{\bV}{{\mathbf V}}
124: \newcommand{\bW}{{\mathbf W}}
125: \newcommand{\bX}{{\mathbf X}}
126: \newcommand{\bY}{{\mathbf Y}}
127: \newcommand{\bZ}{{\mathbf Z}}
128: \newcommand{\asc}{{\bf asc}}
129: 
130: \newcommand{\sA}{{\mathcal A}}
131: \newcommand{\sC}{{\mathcal C}}
132: \newcommand{\sD}{{\mathcal D}}
133: \newcommand{\sG}{{\mathcal G}}
134: \newcommand{\sI}{{\mathcal I}}
135: \newcommand{\sJ}{{\mathcal J}}
136: \newcommand{\sL}{{\mathcal L}}
137: \newcommand{\sM}{{\mathcal M}}
138: \newcommand{\sP}{{\mathcal P}}
139: \newcommand{\sS}{{\mathcal S}}
140: \newcommand{\sX}{{\mathcal X}}
141: \newcommand{\PD}{\sP_{\sD}}
142: \newcommand{\GD}{\sG_{\sD}^0}
143: \newcommand{\bsq}{\vrule height .9ex width .8ex depth -.1ex}
144: 
145: 
146: \makeatletter
147: % put a period after section or subsection number in header
148: \def\@sect#1#2#3#4#5#6[#7]#8{\ifnum #2>\c@secnumdepth
149:      \def\@svsec{}\else
150:      \refstepcounter{#1}\edef\@svsec{\csname the#1\endcsname.\hskip .75em }\fi
151:      \@tempskipa #5\relax
152:       \ifdim \@tempskipa>\z@
153:         \begingroup #6\relax
154:           \@hangfrom{\hskip #3\relax\@svsec}{\interlinepenalty \@M #8\par}%
155:         \endgroup
156:        \csname #1mark\endcsname{#7}\addcontentsline
157:          {toc}{#1}{\ifnum #2>\c@secnumdepth \else
158:                       \protect\numberline{\csname the#1\endcsname}\fi
159:                     #7}\else
160:         \def\@svsechd{#6\hskip #3\@svsec #8\csname #1mark\endcsname
161:                       {#7}\addcontentsline
162:                            {toc}{#1}{\ifnum #2>\c@secnumdepth \else
163:                              \protect\numberline{\csname the#1\endcsname}\fi
164:                        #7}}\fi
165:      \@xsect{#5}}
166: % put a period after theorem and theorem-like numbers
167: \def\@begintheorem#1#2{\it \trivlist \item[\hskip \labelsep{\bf #1\ #2.}]}
168: 
169: \def\plain{plain}\ifx\fmtname\plain\csname fi\endcsname
170:      \def\batchfile{here.doc}
171:      \input docstrip
172:      \preamble
173: 
174:      Do not distribute the stripped version of this file.
175:      The checksum in the header refers to the documented version.
176: 
177:      \endpreamble
178:      \generateFile{here.sty}{t}{\from{here.doc}{}}
179:      \endinput
180: \fi
181: \ifcat a\noexpand @\let\next\relax\else\def\next{%
182:     \documentstyle[here,doc]{article}\MakePercentIgnore}\fi\next
183: \ifx\@Hxfloat\@Hundef\else\expandafter\endinput\fi
184: \let\@Hxfloat\@xfloat
185: \def\@xfloat#1[{\@ifnextchar{H}{\@HHfloat{#1}[}{\@Hxfloat{#1}[}}
186: \def\@HHfloat#1[H]{%
187: \expandafter\let\csname end#1\endcsname\end@Hfloat
188: \vskip\intextsep\vbox\bgroup\def\@captype{#1}\parindent\z@
189: \ignorespaces}
190: \def\end@Hfloat{\egroup\vskip \intextsep}
191: 
192: \makeatother
193: 
194: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
195: 
196: \catcode`\@=11
197: %\renewcommand{\section}{
198: %        \setcounter{equation}{0}
199: %        \@startsection {section}{1}{\z@}{-3.5ex plus -1ex minus
200: %        -.2ex}{2.3ex plus .2ex}{\large\bf}
201: %        }
202: 
203: \catcode`@=12
204: 
205: 
206: \thispagestyle{empty}
207: 
208: \begin{document}
209: 
210: 
211: 
212: % TITLE PAGE and ABSTRACT
213: 
214: \begin{center}
215: {\Large {\bf Apollonian Circle Packings: Geometry and Group Theory}}\\
216: {\Large {\bf II. Super-Apollonian Group and Integral Packings}}\\
217: \vspace{1.5\baselineskip}
218: {\em Ronald L. Graham}\\
219: Department  of Computer Science and Engineering\\
220: University of California at San Diego, 
221: La Jolla, CA 92093-0114 \\
222: \vspace*{1.5\baselineskip}
223: 
224: {\em Jeffrey C. Lagarias} \\
225: Department of Mathematics \\
226: University of Michigan, 
227: Ann Arbor, MI 48109--1109 \\
228: \vspace*{1.5\baselineskip}
229: 
230: {\em Colin L. Mallows} \\
231: Avaya Labs, Basking Ridge, NJ 07920 \\
232: \vspace*{1.5\baselineskip}
233: 
234: {\em Allan R. Wilks} \\
235: AT\&T Labs, Florham Park, NJ 07932-0971 \\
236: \vspace*{1.5\baselineskip}
237: 
238: {\em Catherine H. Yan}
239: \footnote{
240: Partially supported by NSF grants DMS-0070574, DMS-0245526 and a Sloan
241: Fellowship. This author is also affiliated with Dalian University of
242: Technology, China.}\\
243: Department of Mathematics \\
244: Texas A\&M University,
245: College Station, TX 77843-3368\\
246: \vspace*{1.5\baselineskip}
247: %\today
248: (March 10, 2005 version) \\
249: \vspace*{1.5\baselineskip}
250: {\bf ABSTRACT}
251: \end{center}
252: Apollonian circle packings arise by repeatedly filling the interstices
253: between four mutually tangent circles  with further tangent circles.
254: Such packings can be described in terms of the Descartes configurations
255: they contain, where a  
256: Descartes configuration is a set of four mutually tangent
257: circles in the Riemann sphere, having disjoint interiors.
258: Part I showed there exists a discrete group, the Apollonian group,
259: acting on a parameter space of 
260: (ordered, oriented) Descartes configurations, 
261: such that the Descartes configurations in a packing formed an
262: orbit under the action of this group. It is observed there
263: exist infinitely many types of integral
264: Apollonian  packings in which all circles had integer curvatures,
265: with the integral structure being related to the integral
266: nature of the Apollonian group. 
267: Here we consider the
268: action of a larger discrete group, the super-Apollonian
269: group, also having an integral structure,
270: whose orbits describe the Descartes quadruples of a geometric
271: object we call a 
272: super-packing. The circles in a super-packing never cross each other
273: but are nested to an arbitrary depth.
274: Certain Apollonian packings and super-packings
275: are strongly integral in the sense that the curvatures of all
276: circles are integral and the curvature$\times$centers of all
277: circles are integral. We show that (up to scale)
278: there are exactly $8$ different
279: (geometric) strongly integral super-packings, and that each
280: contains a copy of every integral Apollonian circle packing
281: (also up to scale). We show  that the super-Apollonian group
282: has finite volume in the group of all automorphisms of the
283: parameter  space of 
284: Descartes configurations, which is isomorphic
285: to the Lorentz group $O(3, 1)$.
286: 
287: \vspace*{1.5\baselineskip}
288: \noindent
289: Keywords: Circle packings, Apollonian circles, Diophantine equations,
290: Lorentz group, Coxeter group
291: 
292: \setcounter{page}{1}
293: 
294: 
295: %*****************************************************************************
296: %
297: % SECTION 1. 
298: %
299: %*****************************************************************************
300: %
301: \setlength{\baselineskip}{1.0\baselineskip}
302: \section{Introduction}
303: \setcounter{equation}{0}
304: 
305: 
306: Apollonian circle packings are arrangements of
307: tangent circles that arise by repeatedly filling the interstices
308: between four mutually tangent circles  with further tangent circles.
309: A set of four mutually tangent circles is called a
310: Descartes configuration. 
311: Part I studied Apollonian circle packings in terms of
312: the set of Descartes configurations that they contain.
313: It is observed that there exist  Apollonian circle packings 
314: that have a very strong integral structure, with all
315: circles in the packing having integer curvatures,
316: and rational centers, such that curvature$\times$center
317: is an integer vector. We termed these {\em strongly
318: integral} Apollonian circle packings. An example is 
319: the $(0,0,1,1)$ packing pictured in Figure~\ref{fig1},
320: with the two circles of radius $1$ touching at the origin, and with
321: two straight lines parallel to the $x$-axis.
322: 
323: %
324: % Figure 1  (0,0,1,1)
325: %
326: 
327: \begin{figure}[htbp]
328: \centerline{\epsfxsize=5.0in \epsfbox{0.0.1.1.ps}}~\label{fig1}
329: \caption{The integer  Apollonian packing $(0,0, 1,1)$}
330: \end{figure}
331: 
332: 
333: Part I gave an explanation for the existence of such
334: integral structures. This uses  a coordinate
335: system for describing all (ordered, oriented)
336: Descartes configurations $\sD$ in terms of their 
337: curvatures and centers, which forms a $4 \times 3$
338: {\em curvature-center coordinate}
339: matrix $\bM_{\sD}$, and a more detailed coordinate system,
340: {\em augmented curvature-center coordinates},
341: using $4\times 4$ matrices $\bW_{\sD}$. 
342: The strongly integral property of a single Descartes configuration
343: is encoded in the integrality of the matrix $\bM_{\sD}$.
344: The set of all (geometric) Descartes configurations in an 
345: Apollonian packing can be described as a single  orbit of
346: a certain discrete group $\sA$ of $4 \times 4$ 
347: integer matrices; algebraically there are $48$  orbits
348: of ordered, oriented Descartes configurations giving rise to 
349: the same geometric packing, which  correspond to the 
350: 48 possible ways of ordering the circles and totally orienting the
351: configuration.
352: The integrality of the members
353: of $\sA$ is the source of the strong integrality
354: of some Apollonian circle packings. 
355: As a consequence of this group action, if  a single
356: Descartes configuration in the packing is strongly
357: integral, then they all are; hence  
358: every individual circle in the
359: packing is strongly integral. 
360: 
361: There are infinitely many distinct integral Apollonian
362: circle packings. Two more  of them are pictured in
363: Figures 2 and 3,  the $(-1, 2,2,3)$-packing
364: and the $(-6, 11, 14, 15)$-packing, respectively.
365:  Any integral packing
366: can be moved by a Euclidean motion so as to be
367: strongly integral, as will follow from results in
368: this paper.
369: 
370: 
371: %
372: % Figure 2  (-1,2,2,3)
373: %
374: 
375: \begin{figure}[htbp]
376: \centerline{\epsfxsize=3.0in \epsfbox{1.2.2.3.ps}}
377: \caption{The integer  Apollonian packing $(-1,2,2,3)$}~\label{fig2}
378: \end{figure}
379: %
380: % Figure 3 (-6, 11, 14, 15))
381: %
382: %
383: \begin{figure}[htbp]
384: \centerline{\epsfxsize=3.0in \epsfbox{6.11.14.15.ps}}
385: \caption{The integer  Apollonian packing $(-6,10,11,14)$}~\label{fig3}
386: \end{figure}
387: 
388: 
389: Part I introduced some further integral actions on
390: Descartes configurations, involving a ``duality''
391: operator $\bD$ leading to a dual Apollonian group $\sA^{\perp}$.
392: Combining this group with the Apollonian group
393: leads to a large group of integer $4 \times 4$ matrices,
394: the {\em super-Apollonian group} $\sA^{S}$, which 
395: acts on the set of all (ordered, oriented) Descartes
396: configurations. Part I showed that the super-Apollonian
397: group is a hyperbolic Coxeter group. 
398: It defined an {\em Apollonian super-packing}
399: to be an orbit  $\sA^{S}[\sD]$ of a single
400: Descartes configuration $\sD$ under this group. 
401: Such a super-packing is called  {\em integral}
402: if the initial Descartes configuration has integer
403: curvatures, and is  
404: {\em strongly integral} if the curvature-center coordinates
405:  $\bM_{\sD}$ of the initial Descartes configuration
406: are integral. These properties hold for all Descartes
407: configurations in the packing if they hold for one.
408: 
409:  In this paper we study the geometric 
410: structure  of super-packings,
411: and their integrality properties.
412: A  striking geometric fact is
413:  that the circles in a super-packing do not
414: overlap, as shown in \S3. Figure~\ref{fig4}  shows
415: circles of curvature at most $65$ in the 
416: super-packing generated from the $(0,0,1,1)$ 
417: configuration in Figure~\ref{fig1} above.
418: (The generating Descartes configuration is indicated
419: with slightly darker lines.)
420: Here the horizontal and vertical scales of the
421: figure are roughly from $-2.2 \le x \le 2.2$
422: and $-2.2 \le y \le 2.2$. This picture is 
423: representative of  the
424: whole super-packing; we show in \S6 that this
425: super-packing is periodic under the 
426: two-dimensional lattice generated
427: by the shifts
428: $x \to x+2$ and $y \to y+2$.  
429: 
430: %
431: % Figure 4  (superpacking to depth 20)
432: %
433: %
434: \begin{figure}[htbp]
435: \centerline{\epsfxsize=6.0in \epsfbox{super.ps}}
436: \caption{An initial part of the  $(0,0,1,1)$ super-packing
437: (square of sidelength $4.4$)}~\label{fig4}
438: \end{figure}
439: 
440: We cannot picture the whole super-packing because
441: the circles in it are dense in the plane; they nest inside
442: each other to an arbitrary depth. One can show that
443: the set of points lying in the interior of an infinite
444: nested sequence of such circles has full Lebesgue
445: measure. However there is interesting structure
446: visible if we turn up the scale of the circles.
447: In \S6 in Figure~\ref{fig8} we show all circles of curvature at
448: most 200 inside the super-packing above, inside
449: the unit square $0 \le x \le 1,~ 0 \le y \le 1$.
450: 
451: 
452: This paper is concerned with integrality 
453: properties of some super-packings, and we study successively
454: stronger integrality properties. In \S4 we require the
455: curvatures to be integral, in \S5 and \S6 both
456: the curvatures and curvature$\times$center must be 
457: integral, and in \S8 the augmented curvature-center
458: coordinates of all Descartes configurations in the 
459: super-packing must be integral. At this last level we
460: show that there are exactly $14$ such super-packings, viewed
461: as rigid geometric objects.
462: In \S2 we give a more detailed description of our results.
463: 
464: The general framework of this paper 
465: was developed by  the second author (JCL), who also did much of
466: the writing. This paper is an extensive
467: revision  of a preprint written in 2000, and adds new
468: results in \S3 and  \S6.
469: 
470: 
471: \paragraph {Acknowledgments.} The authors are grateful for
472: helpful comments from Andrew Odlyzko, Eric Rains,
473: Jim Reeds and  Neil Sloane during this work. They
474: also thank the  reviewer for incisive comments. \\
475: 
476: %*****************************************************************************
477: %
478: % SECTION 2. Summary
479: %
480: %*****************************************************************************
481: %
482: \section{Summary of Main Results}
483: \setcounter{equation}{0}
484: 
485: We consider Apollonian super-packings.
486: Analogously to part I,
487: an   Apollonian  super-packing may be considered as either
488: an geometric object or an algebraic object, as follows. \\
489: 
490: (i) [Geometric] A {\em geometric Apollonian super-packing} 
491: is a point set on the Riemann sphere
492: $\hat{\cc} = \RR^2 \cup \{ \infty\}$, consisting of 
493: all the circles in four orbits
494: of a certain group $G_{\sA^{S}}(\sD)$ of
495: M\"{o}bius transformations
496: inside  the conformal group M\"{o}b(2) acting
497: on the four circles $\{C_1, C_2, C_3, C_4\}$ in
498: a given Descartes configuration $\sD$.  The 
499: group $G_{\sA^{S}}(\sD)$ depends on $\sD$. \\
500: 
501: (ii) [Algebraic] An {\em (algebraic) Apollonian super-packing} is a
502:  set of ordered, oriented Descartes configurations,
503: given by $48$ orbits of the
504: super-Apollonian group $\sA^{S}[\sD]$. The augmented 
505: curvature-center coordinates of
506: its elements are $\sA^{S}\bW_{\sD}:=\{ \bU W_{\sD} : \bU \in \sA^{S}\}$. \\
507: 
508: 
509: 
510: A geometric super-packing  can be 
511: described  in terms of its unordered, unoriented
512: Descartes configurations. From this viewpoint,
513:  each geometric Apollonian
514: super-packing corresponds to  $48$ different algebraic 
515: super-packings; there are 
516: $24$ choices of ordering the four circles 
517: and $2$ choices of total orientation of the configuration. 
518: We can  consider geometric  super-packings as unions
519: of a countable  number of Apollonian packings. 
520: This leads to interesting questions concerning the way these
521: Apollonian packings are embedded inside the geometric super-packing.
522: We note that as  a point set, the geometric super-packing is
523: invariant under the group action $G_{\sA^{S}}(\sD)$. 
524: However it is not a closed set,   and its closure is
525: the entire Riemann sphere $\RR^2 \cup \{ \infty\}$.
526: 
527: A large part of the paper considers 
528: integrality properties of curvatures and centers
529: of some super-packings. These questions are mainly
530: studied using algebraic super-packings, although
531: we also consider questions concerning 
532: the associated geometric super-packing, such as
533: its group of symmetries under Euclidean motions.
534: 
535: In \S3  Theorem~\ref{th31} shows that each  geometric 
536: super-packing is 
537: a packing in the sense that the circles in it do not cross
538: each other transversally, as mentioned above.
539: Theorem~\ref{th32n} specifies certain subcollections of 
540: geometric super-packings which are genuine packings
541: in the sense that the interiors of the circles do not
542: overlap.
543: 
544: In \S4 we study integer super-packings,
545: in which all circles have integer curvatures.
546: Integer super-packings are classified up to
547: Euclidean motions by a 
548: single invariant, their {\em divisor} $g$,
549: which is the greatest common divisor of the
550: curvatures in any Descartes configuration in
551: the packing. Theorem~\ref{th41n} shows  that for
552: each $g \ge 1$ there is a unique such
553: integral super-packing, up to a Euclidean
554: motion. We also show in Theorem~\ref{th43n} 
555: that for each geometric Apollonian circle 
556: packing that is integral, there exists a
557: Euclidean motion taking it to one that is
558: strongly integral.
559: 
560: In \S5 we study strongly integral super-packings,
561: which are those whose curvatures are integral and whose
562: curvature$\times$center is also integral.
563: Strongly-integral super-packings are geometrically rigid:
564: Theorem~\ref{orbits} shows that for each integer $g \ge 1$ there are
565: exactly $8$ strongly integral geometric super-packings
566: which have divisor $g$. Here we do not allow the packings
567: to be moved by Euclidean motions.
568: 
569: In \S6 we study the relations between integer Apollonian packings
570: and strongly integral super-packings. 
571: Without loss of generality we restrict to
572: primitive integer super-packings, those with divisor $1$.
573: Theorem~\ref{th61} shows
574:  that each of the $8$ kinds of these has a large group of
575: internal symmetries, forming a crystallographic group.
576: For convenience we fix one of them and call it the
577: {\em standard strongly integral super-packing}; results proved for
578: it have analogues for the other seven. Theorem~\ref{th62}
579: shows that each primitive integral Descartes configuration
580: (except for the $(0,0,1,1)$ 
581: configuration) occurs in this packing with the center of its largest
582: circle being contained in the closed unit square
583: $0 \le x \le 1, ~ 0 \le y \le 1$, 
584: and the location of this circle center is unique.  
585: Theorem~\ref{th63} deduces  that the geometric 
586: standard strongly integral  super-packing contains a
587: unique copy of each primitive integral Apollonian packing, 
588: except for  the
589: $(0,0,1,1)$ packing, 
590: having the property that 
591: the center of its bounding circle lies inside the
592: closed unit square.
593: Figures~\ref{fig5}, \ref{fig6} and \ref{fig7} picture
594: the locations of all primitive integer Apollonian
595: packing with bounding circle of curvatures 6, 8 and
596: 9, respectively. The unit square is indicated by 
597: slightly darker shading in the figures. 
598: Note that in Figure~\ref{fig5} the three Apollonian packings
599: are generated by Descartes configurations with
600: curvature vectors $(-6, 7, 42, 43),~ (-6, 10, 15,19)$
601: and $(-6,11, 14, 15)$; these are root quadruples in 
602: the sense of \cite[Sec. 4]{GLMWY2}.
603: 
604: %
605: % Figure 5  (N=- 6 configurations)
606: %
607: %
608: \begin{figure}[htbp]
609: \centerline{\epsfxsize=3.0in \epsfbox{all6.ps}}
610: \caption{Integer 
611: Apollonian packings with bounding circle of curvature $6$.}~\label{fig5}
612: \end{figure}
613:  
614: %
615: % Figure 6  (N= -8  configurations)
616: %
617: %
618: \begin{figure}[htbp]
619: \centerline{\epsfxsize=3.0in \epsfbox{all8.ps}}
620: \caption{Integer Apollonian packings with bounding circle of curvature $8$.
621: }~\label{fig6}
622: \end{figure}
623:  
624: %
625: % Figure 7  (N=-9 configurations)
626: %
627: %
628: \begin{figure}[htbp]
629: \centerline{\epsfxsize=3.0in \epsfbox{all9.ps}}
630: \caption{Integer Apollonian packings with bounding circle of curvature $9$.
631: }~\label{fig7}
632: \end{figure}
633: 
634: 
635: 
636: In \S7 Theorem~\ref{th51} shows  that the super-Apollonian
637: group is a finite index normal subgroup of the 
638: group $Aut(Q_{D}, \ZZ)$ of integral automorphs of 
639: the Descartes quadratic form. The latter group can
640: be identified with an index $2$ subgroup of the
641: integer
642: Lorentz group $O(3, 1, \zz)$, and this identification
643: allows us to identify 
644: the super-Apollonian group with  a 
645: particular normal subgroup $\tilde{\sA}^{S}$
646: of index $96$ in  $O(3, 1, \zz)$, defined after
647:  Theorem~\ref{th51}.
648: 
649: In  \S8 we study  super-packings all of whose 
650: Descartes configurations are {\em super-integral} in the
651: sense that their augmented curvature-center coordinate
652: matrices $\bW_{\sD}$ are integer matrices. Theorem~\ref{th81n} 
653: shows  there
654: are exactly $14$ geometric super-packings of this kind.
655: 
656: The last section \S9 makes a few concluding remarks.
657: 
658: The Appendix considers minimal conditions to guarantee
659: that a Descartes configuration
660: is strongly integral.  Theorem~\ref{th101} shows
661: that a configuration is strongly integral if (and only
662: if) three of its four circles are strongly integral. 
663: 
664: 
665: 
666: 
667: 
668: 
669: %*****************************************************************************
670: %
671: % SECTION 3. Geometric super-packings
672: %
673: %*****************************************************************************
674: %
675: \section{Geometric Apollonian Super-Packings}
676: \setcounter{equation}{0}
677: 
678: 
679: In this section we consider properties of 
680: geometric Apollonian circle packings. We view such
681: a super-packing as a point set on the Riemann sphere
682: $\hat{\cc} = \cc \cup \{\infty\}.$  We first
683: note  that it is not a closed set.
684: It  is not hard to show that its  closure is
685: the whole Riemann sphere, 
686: Each geometric Apollonian packing  has  a group invariance
687: property under a certain group of M\"{o}bius transformations
688: which depends on the super-packing, which is 
689: the group generated by the countable set of  
690: groups of M\"{o}bius transformations that leave 
691: invariant some Apollonian packing contained in
692: the super-packing. 
693: 
694: Our object is  prove the following
695: ``packing'' property of geometric super-packings.
696: %
697: %
698: % Theorem 3.1
699: %
700: %
701: 
702: \begin{theorem}~\label{th31} 
703: A  geometric Apollonian super-packing is a circle
704: packing in the weak sense that no two circles belonging
705: to it cross each other transversally. Circles in the geometric 
706: super-packing may be nested, or tangent to each other.
707: \end{theorem}
708: 
709: 
710: Before giving the proof, we describe the 
711: nature of the geometric packing in terms of 
712: nesting of circles. We view the packing 
713: $\sA^{S}[\sD_0]$ as generated from an initial
714: (positively oriented) Descartes configuration $\sD_0$,
715: by multiplication by a finite set of generators of
716: the super-Apollonian group.
717: Each  circle in
718: the super-packing has a  well-defined 
719: {\em nesting depth} $d$ 
720: (relative to the generating configuration $\sD_0$)
721: which counts the number of circles in the packing which
722: include $C$ in their interior. Here the notion of
723: ``interior'' is defined with respect to  the initial Descartes
724: configuration $\sD_0$.
725: The Apollonian group
726: generators move ``horizontally'', leaving constant
727: the nesting depth of any circles they produce. 
728: The dual Apollonian group generators move ``vertically'',
729: by reflecting three of the circles in a configuration
730: into the interior of the fourth circle, they increase
731: the nesting depth by one. We show there is a unique
732: ``normal form'' word of minimal length in the generators
733: that produces a Descartes
734: configuration $\sD$ containing $C$.
735: The nesting depth of $C$ 
736: is exactly equal to the number $d$ of generators
737: of $\sA^{\perp}$ that appear in this normal form word.
738: The circles at nesting depth $0$
739: are those circles in the Apollonian packing generated by
740: $\sD_0$. Each circle $C$ at nesting depth $k \ge 1$ contains
741: a unique Apollonian packing, consisting  of it
742: plus all circles at depth $k+1$ nested inside it.
743: 
744: 
745: \paragraph{Proof of Theorem~\ref{th31}.} 
746: We view the geometric super-packing on the Riemann sphere
747: $\hat{\cc}= \cc \cup \{ \infty \},$ so the initial 
748: Descartes configuration $\sD_0$ 
749: consists of four circles on the sphere. In this case each
750: circle defines a spherical cap. We choose an ordering
751: and orientation of $\sD_0$ (this does not affect the geometry),
752: requiring that  $\sD_0$ have  positive (total)  orientation. 
753: 
754: 
755: Let the super-packing be $\sA^{S}[\sD_0] = \sA^{S} \bW_{\sD_0}.$
756: We  consider the effect of the super-Apollonian group generators
757: acting on the left on
758: the matrix $\bW_{\sD_0}$.
759: The Descartes
760: configurations in the super-packing are given by 
761: $\sD = \bU_m \bU_{m-1} \cdots \bU_1[\sD_0]$
762: with
763: each $\bU_k \in \{ \bS_1, \bS_2, \bS_3, \bS_4,
764: \bS_1^{\perp}, \bS_2^{\perp}, \bS_3^{\perp}, \bS_4^{\perp}\}$.
765: We consider words in the group measured by their {\em length} $m$.
766: The stage $m$ circles will consist of all new circles added
767: using products of $m$ generators. 
768: We may without loss of generality restrict to 
769: {\em normal form} words, which are those that satisfy
770: the two conditions:
771: 
772: (i) If $\bU_k = \bS_i$, then $\bU_{k+1} \ne \bS_i$
773: and $\bU_{k+1} \ne \bS_j^{\perp}$ with $j \ne i$.
774: 
775: (ii) If $\bU_k = \bS_i^{\perp}$, then  $\bU_{k+1} \ne \bS_i^{\perp}$. 
776: 
777: Equivalently, looking backwards, if $\bU_{k+1} = \bS_i^{\perp}$
778: then either $\bU_k= \bS_i$ or else $\bU_{k}=\bS_j^{\perp}$ 
779: for some $j \ne i$. A word may be put in
780: normal form by canceling adjacent equal generators, since
781: all $\bS_i^2 = (\bS_i^{\perp})^2= \bI$, and by moving towards
782: the right\footnote{That is, moving it towards
783: the beginning of the word.}
784: in the word as far as possible any generator $\bS_i^{\perp}$,
785: using the property that it commutes with all $\bS_j$ with  $j \ne i$. 
786: These operations eventually put a word in normal form,
787: without increasing its length. The operations do not change
788: the Descartes configuration $\sD$ it represents.
789: 
790: We prove the theorem by induction on the
791: number of symbols $m$ in a normal form word,
792: which we call the stage of the induction.
793: The induction hypotheses at stage $m$ are  as follows.
794: Here we  let $C_i$ refer to
795: the circle at row $i$ of the associated
796: (ordered, oriented) Descartes configuration.
797: 
798: 
799: (1) Each normal form word of length $m$ produces
800: either one or three new circles, according as
801: $\bU_{m}=\bS_i$, where it is the circle
802: $C_i$ or $\bU_{m}= \bS_j^{\perp}$, where it is
803: the three circles $C_i$ with $i \ne j$. 
804: 
805: (2) The nesting depth of any new circle produced
806: at this stage is equal to the number of occurrences of a letter
807: in $\sA^{\perp}$ in its generating normal form
808: word.
809: 
810: (3) Each new circle produced has empty interior,
811: when it first appears. 
812: 
813: In particular hypothesis (3) implies that all circles
814: produced at stage $m$ have disjoint interiors, and 
815: that each such circle  
816: contains no circles from  earlier stages in 
817: its interior. 
818: If the induction is proved, then 
819: hypothesis  (3) guarantees  that the nesting depth of a
820: circle is well-defined when it is first produced,
821: because no new circle will ever include it in its
822: interior. Hypothesis  (3) also  guarantees  that all
823: circles produced by the end of stage $m+1$ do
824: not cross. As a consequence  no two circles in
825: the packing cross, using  property (3) 
826: applied at that level $m$ which is 
827: the greater of the levels of the two circles.
828: Thus the theorem will follow.
829: 
830: The base case $m=0$ of the
831: induction is immediate, consisting of the
832: initial Descartes configuration $\sD_0$.
833: We now show the induction step for $m+1$, given $m$. 
834: We are given  a Descartes configuration
835: $\sD' = \bU_{m+1} [\sD] = \bU_{m+1} \bU_m \cdots \bU_1[\sD_0]$
836: with a normal form word. 
837: 
838: To establish hypothesis (1) for $m+1$, suppose first
839: that  $\bU_{m+1} = \bS_i^{\perp}.$
840: We assert that  the $i$-th circle $C_i$ of $\sD$ was a new
841: circle produced at stage $m$. For either $\bU_m = \bS_i$,
842: in which case it was the unique new circle in $\sD$
843: by induction hypothesis (1) at stage $m$, or $\bU_m = \bS_j^{\perp}$
844: with $j \ne i$, in which case it was one of three new circles
845: produced at stage $m$. By induction hypothesis (2) the
846: circle $C_i$ has empty interior at stage $m$. The
847: three circles $\{C_j^{'}: j \ne i\}$ in 
848: the new configuration $\sD'$  are contained
849: in the interior of  $C_i$ are therefore new circles.
850: They do not cross, being part of a Descartes configuration.
851: Thus hypothesis (1) holds for $m+1$ in this case. 
852: 
853: Suppose next that $\bU_{m+1} = \bS_i$,
854: so the possible new circle is $C_i^{'}$.
855: If 
856: $\sD_k:=\bU_k \bU_{k-1} \cdots \bU_1[\sD_0]$ 
857: is  the maximal 
858: length subword
859: such that $\bU_k= \bS_j^{\perp}$ for some $j$, 
860: with $1 \le k \le m$ then 
861: $\sD'$ belongs to the Apollonian packing generated by
862: $\sD_k$, since all subsequent generators belong to the
863: Apollonian group. If no such $k$ exists, then
864: $\sD'$ is in the Apollonian packing generated
865: by the original Descartes configuration $\sD_0$.
866: For $k \ge 1$, this  Apollonian packing is
867: entirely contained  in
868: the interior of a 
869: bounding circle $C= C_j$ first produced
870: at stage $k-1$. At that time the interior of $C$ was empty,
871: by induction hypotheses (2) and (3) at stage $k-1$.
872: The only Descartes configurations that ever can enter
873: the interior of the circle $C$, must do so by
874: a reflection in $C$ and these   are exactly
875: those normal form words
876: starting with initial segment $\bU_k \bU_{k-1} \cdots \bU_1$.
877: (This follows from uniqueness of a circle when it is
878: created, hypothesis (3) applied at stage $k$.)
879: The words contain this initial segment at the same depth,
880: with all subsequent letters in $\sA$, fill out an Apollonian
881: packing at this depth.
882:  In particular each such normal form word produces one
883: new circle in this Apollonian packing.
884: Recall that  all the circles
885: in the Apollonian packing inside the bounding circle $C$
886: have disjoint interiors (Theorem 4.1 of part I). 
887: These circles all have the same depth, and (1) holds in
888: this case. Normal form words that have another subsequent
889: generator in $\sA^{\perp}$ confine the resulting 
890: Descartes configuration to the
891: inside  of a  single circle 
892: in the Apollonian packing $\sA[\sD_k]$ already produced,
893: and all longer words with this prefix are entirely
894: contained inside this circle. In particular,
895: they do not coincide with  $C_i^{'}$.
896: the new circle produced by the normal form word corresponding
897: to $\sD'$.
898: %(and are disjoint from its interior).
899: It follows that the circle $C_i^{'}$ is new,
900: so hypothesis (1) holds for $m+1$ in this case.
901: 
902: There remains the case where $\bU_{m+1} = \bS_i$
903: and all previous $\bU_k \in \sA$. Then 
904: $\bU_{m+1} \cdots \bU_1[\sD_0]$ belongs
905: to the Apollonian packing generated by
906: $\sD_0$, and is at depth $0$. Here $C_i^{'}$
907: is new since the generation of an 
908: Apollonian packing creates one new circle
909: at each step, and any word that contains
910: an element of $\sA^{\perp}$ moves the
911: Descartes configuration inside an older
912: circle in this packing, from which it cannot
913: escape. Thus $C_i^{'}$ is a new circle in
914: this case, and hypothesis (1) holds.
915: 
916: 
917: 
918: Hypothesis (2) holds for $m+1$ in the case $\bU_m =\bS_i^{\perp}$
919: because the three new circles produced have nesting
920: depth one greater than the nesting depth of $C_i$ at
921: the previous level, to which induction hypothesis (2) applied.
922: It also holds in the remaining case $\bU_m=\bS_i$ because the 
923: argument above showed that the nesting 
924: depth did not increase in this case.
925: 
926: Hypothesis (3) holds if $\bU_{m+1} =\bS_i^{\perp}$. Now 
927: circle $C_i$ was first created only at stage $m$,
928: and the only other possible sequences leading to
929: a Descartes configuration including 
930: this circle 
931: must start from $\sD_m$ and use a generator
932: $\bU_{m+1}= \bS_j$
933: with  $j \ne i$.  In all other cases the resulting
934: Descartes configuration includes no circle inside
935: $C_i$, so the interiors of the three new circles 
936: produced are empty at the end of stage $m+1$.
937: 
938: In the remaining case  $\bU_{m+1} =\bS_i$, we have
939: already observed that the new circle $C_i^{'}$ produced,
940: is disjoint from all other circles in the Apollonian
941: packing  $\sA[\sD_k]$ created by level $m$, 
942: which are necessarily contained  in the
943: bounding circle $C$ of $\sD_k$. 
944: As mentioned above, all Descartes configurations
945: containing a circle inside $C$ must have an
946: initial segment of their generating word, giving
947: the unique normal form word that first generates $C$,
948: at stage $k \le m$. Other depth $m+1$ words with 
949: this initial segment, and with all subsequent $\bU_j$ drawn 
950: from the Apollonian group generators 
951: produce new circles in the
952: Apollonian packing $\sA[\sD_k]$, disjoint from $C_i^{'}$.
953: Any  normal  form word
954: at depth $m+1$ with this initial segment, which 
955: contain  some generator $\bS_j^{\perp}$ after-wards,
956: produce a Descartes configuration contained inside a circle
957: of the Apollonian packing $\sA[\sD_k]$
958: different from $C$. 
959: It follows that the
960: interior of $C_i^{'}$ is empty at the end of stage
961: $m+1$, as required.
962: 
963: This completes the induction step and the proof.
964: $~~~\bsq$ \\
965: 
966: Super-packings have  some properties that
967: are genuine packing properties. The proof of Theorem~\ref{th31}
968: established in hypothesis  (3) shows that the finite set of circles at
969: stage $m$ of the construction, starting from a generating
970: Descartes configuration $\sD$,  all had disjoint interiors.
971: It also gives  the following stronger result. 
972: 
973: %
974: %
975: % Theorem 3.2
976: %
977: %
978: 
979: 
980: \begin{theorem}~\label{th32n} 
981: For a geometric Apollonian super-packing given with
982: a generating Descartes configuration $\sD$, 
983: and each $k \ge 1$,  
984: the set of all circles having  nesting depth exactly $k$ 
985: with respect to $\sD$ have pairwise  disjoint interiors.
986: These circles can be viewed as  
987: forming an infinite collection
988: of Apollonian packings, each missing one circle;
989: the missing circle is a bounding circle at depth $k-1$.
990: \end{theorem}
991: 
992: \paragraph{Proof.} 
993: The proof of Theorem~\ref{th31}
994: shows that the nesting depth of a
995: circle is well-defined.  Circles at nesting depth $k$ have
996: disjoint interiors, since no two of them cross, and the
997: only way for two of them to have an interior point in
998: common is for one to be nested inside the other, which 
999: would violate the nesting ordering. 
1000: 
1001: Given a normal form word $\bU := \bU_m \bU_{m-1} \cdots \bU_1$
1002: with $\bU_m = \bS_i^{\perp}$, and containing exactly
1003: $k$ elements of 
1004: $\sA^{\perp} = 
1005: \langle \bS_1^{\perp},\bS_2^{\perp}, \bS_3^{\perp}, \bS_4^{\perp} \rangle, $
1006: the set of all normal form words having $\bV$ as prefix and all
1007: other letters $\bU_j \in \sA$ for $j > m$ produce all the
1008: circles in an Apollonian packing $\sA[\sD_m]$, 
1009: with $\sD_m = \bU[\sD_0]$.
1010: All these circles are  at depth $k$ 
1011: except for  the outer circle of $\sD_m$, which 
1012: is its $i$-th circle.
1013: Enumerating all possible such $\bU$ as prefixes  represents
1014: the set of nesting depth $k$ circles as a collection of
1015: Apollonian packings, each excluding one circle when $k \ge 1$.
1016: $~~~\bsq$ \\
1017: 
1018: \paragraph{Remarks.}
1019: (1) The proof of Theorem~\ref{th31} is a geometric analogue of the
1020: presentation for the super-Apollonian group
1021: proved in part I, \cite[Theorem 6.1]{GLMWY11}.
1022: 
1023: (2) The analogous result to Theorem~\ref{th31}
1024: fails 
1025: to hold in all dimensions $n \ge 4$, as explained
1026: in part III. The ``nesting'' property
1027: of the dual Apollonian configurations resulting
1028: from ``vertical'' moves still exists
1029: and works in all dimensions.
1030: However the ``horizontal'' motions  moving spheres  around
1031: in Apollonian packings produces  spheres that cross
1032: in dimensions $n \ge 4$, see \cite[Lemma 4.1]{GLMWY13}.
1033: 
1034: 
1035: (3) We cannot easily visualize a geometric super-packing as
1036: a completed object because the circles in it are dense
1037: in the plane. We can however picture a partial version
1038: of it that pictures all circles of size above a given
1039: threshold, in some finite region of the packing.
1040: The integral super-packings we are most interested
1041: in have a periodic lattice of symmetries (see Theorem~\ref{th61}),
1042: so it suffices to examine a finite region of the packing.
1043: Figure~\ref{fig4} in \S1 and Figure~\ref{fig8} in \S6
1044: exhibit part of a super-packing. 
1045: 
1046: (4) Every circle $C$ in a geometric super-packing $\sA^{S}[\sD]$ 
1047: has associated
1048: to it a unique Apollonian packing of which it is the bounding
1049: circle. If  it is a depth $k$ circle (relative  to the
1050: starting configuration $\sD$), then this Apollonian packing 
1051: consists of all depth $k+1$ circles contained in the interior of
1052: $C$. 
1053: 
1054: 
1055: %*****************************************************************************
1056: %
1057: % SECTION 4. Integral  Super-Packings
1058: %
1059: %*****************************************************************************
1060: 
1061: \section{Integral Super-Packings}
1062: \setcounter{equation}{0} 
1063: 
1064: An Apollonian super-packing is {\em integral} if it
1065: contains one (and hence all) Descartes configuration
1066: whose circles have integer curvatures.
1067: 
1068: An invariant of an integral super-packing is its
1069: {\em divisor} $g$, which is the greatest common divisor
1070: of the curvatures of the circles in any Descartes
1071: configuration in the super-packing.
1072: The quantity  $g$ is well-defined   independent of
1073: the Descartes configuration chosen, using the
1074: relation $\bM_{\sD'} = \bU \bM_{\sD}$ between two
1075: such configurations, where $\bU \in \sA^{S}$. Since $\bU$
1076: is an integer matrix with determinant $\pm 1$, it preserves
1077: the greatest common divisor of each column of the integer
1078: matrix $\bM_{\sD}$. Here  the first column
1079: encodes the curvatures of the circle. 
1080: 
1081: %
1082: %
1083: % Theorem 4.1
1084: %
1085: %
1086: 
1087: \begin{theorem}~\label{th41n} 
1088: For each integer $g \ge 1$ there exists an
1089: integral Apollonian super-packing with 
1090: divisor $g$. The associated geometric super-packing
1091: is unique, up to a Euclidean motion.
1092: \end{theorem}
1093: 
1094: As an immediate corollary of this result, the
1095: integral super-packing with divisor $g$ contains
1096: at least one  copy of every integral Descartes configuration
1097: having divisor $g$. For each such a Descartes configuration
1098: generates an integral super-packing with divisor $g$,
1099: so the corollary follows by the uniqueness assertion.
1100: 
1101: We defer the proof of 
1102: Theorem \ref{th41n} to the end of the section.
1103: It is based on a reduction theory which finds inside
1104: any such integral super-packing a Descartes configuration
1105: having particularly simple curvatures. 
1106: %
1107: %
1108: % Theorem 4.2
1109: %
1110: %
1111: 
1112: \begin{theorem}~\label{reduction}
1113: Let $\sD$ be an integral Descartes configuration, with
1114: divisor $g := \gcd(b_1, b_2, b_3, b_4)$.
1115: Then the Apollonian super-packing $\sA^{S} [\sD]$
1116: generated by $\sD$ contains a Descartes configuration having
1117: curvature vector a permutation of either $(0,0, g, g)$
1118: or $(0,0, -g,-g)$, with the former case occurring if
1119: $b_1+b_2+b_3+b_4 > 0$ and the latter case if $b_1+b_2+b_3+b_4 < 0$. 
1120: \end{theorem}
1121: 
1122: \noindent {\bf Proof.} 
1123: Since the super-Apollonian group $\sA^S$ preserves the (total) orientation 
1124: of Descartes configurations, it is sufficient to show that for a positively 
1125: oriented integral Descartes configuration $\sD$ with curvatures
1126: $(b_1, b_2, b_3, b_4)$, $b_1+b_2+b_3+b_4 > 0$, there exists $\bU \in \sA^S$ 
1127: and a permutation matrix $\bP_\sigma$ such that
1128: $$\bP_\sigma \bU(b_1, b_2, b_3, b_4)^T=(0,0,g,g)^T.$$
1129: 
1130: 
1131: We measure
1132: the {\em size} of the curvature vector $\bv = (b_1, b_2, b_3, b_4)^T$
1133: of a  Descartes configuration  by
1134: $$ size(\bv ):= \bo^T \bv = b_1 + b_2 + b_3 + b_4.$$
1135: We claim that for positively oriented integral Descartes configurations
1136: with greatest common divisor $g$, we have
1137: $$size(\bv)\geq 2g,$$
1138: and equality holds if and only if $\bv$ is a permutation of $(0,0,g,g)$.
1139: If all curvatures are nonnegative this is clear, since at most two
1140: can be zero, and the other two are positive integers.
1141: Now in any Descartes configuration  at most one circle can have
1142: negative curvature, call it $b_1=-a$  ($a \in \ZZ_+$), in which case  
1143: it encloses the other three.
1144: Each of these three enclosed circles has a larger curvature in
1145: absolute value than
1146: the bounding circle, so  $b_i \geq a + 1$ for $i=2,3,4$. Thus
1147: $size(\bv) \geq -a+ 3(a + 1) \geq 2a+3 > 2g,$
1148: which proves the claim.
1149: 
1150: We give a reduction procedure which
1151: chooses  matrices in $ \sA^{S}$
1152: to reduce the size and show that the procedure halts only at a
1153: vector of form $(0, 0, g,g)$, up to a permutation.
1154: To specify it, we observe that for the curvature vector 
1155: $\bv=(b_1, b_2, b_3, b_4)^T$ of 
1156: any integral
1157: Descartes configuration  with
1158: $b_1 \leq b_2\leq b_3\leq b_4$, we have
1159: \beql{605aa}
1160: size(\bS_4 \bv) = \bo^T \bS_4\bv \leq \bo^T \bv = size(\bv).
1161: \eeq
1162: and equality holds
1163: if and only if $b_1b_2+b_2b_3+b_3b_1=0$. To see this,
1164: we have $\bS_4\bv=(b_1, b_2, b_3, b_4')^T$ where
1165: $$b_4'=2(b_1+b_2+b_3)-b_4 =b_1+b_2+b_3-2\sqrt{b_1b_2+b_2b_3+b_3b_1}.$$
1166: Thus
1167: $$\bo^T \bS_4 \bv-\bo^T \bv=b_4'-b_4=-4\sqrt{b_1b_2+b_2b_3+b_3b_1} \leq 0.$$
1168: Equality can hold if and only if $b_1b_2+b_2b_3+b_3b_1=0$, which proves the
1169: observation. Note that $b_1 \leq 0$ when the equality in \eqn{605aa} holds.
1170: Also note that $g=\gcd(b_1, b_2, b_3, b_4)$ is an invariant under the 
1171: action of $\sA^S$. 
1172: 
1173: Starting with any  positively oriented integral Descartes configuration with
1174: curvature vector $(b_1, b_2, b_3, b_4)^T$, 
1175: where $b_i$ is the largest number, we apply
1176: $\bS_i \in \sA$.
1177: By \eqn{605aa},  the $size(\bv)$ decreases but cannot be negative,
1178: so after a finite
1179: series of $\bS_i$ we arrive at a positively oriented 
1180: integral Descartes configuration with curvatures 
1181: $\bv'= (b_1', b_2', b_3', b_4')^T$,
1182: where $\gcd(b_1', b_2', b_3', b_4')=g$ and
1183: the smallest curvature, say $b_1'$, satisfies $b_1' \leq 0$, and 
1184: the size of $\bv'$ can not be reduced by the action of the Apollonian group
1185: $\sA$.
1186: Call this the basic reduction step. Note that the basic reduction step
1187: involves only matrices in the Apollonian group and therefore
1188: moves around inside a single Apollonian packing.
1189: 
1190: If $b_1' =0$ then necessarily $b_2'=0$,
1191: whence the curvature vector is $(0,0,b_3',b_3')^T$,
1192: and by $g=\gcd(b_1',b_2', b_3', b_4')$, we have 
1193: $b_3' = g$ and the reduction halts.
1194: If $b_1'<0$, applying $\bS_1^T$, we get a new Descartes configuration with
1195: $\bv''= (-b_1', b_2'+2b_1', b_3'+2b_1', b_4'+2b_1')^T$,
1196: which is positively oriented and lies in a new Apollonian packing
1197: and has $$size(\bv'') = size(\bv) + 4 b_1' < size(\bv).$$
1198: Thus the size strictly
1199: decreases and is non-negative. Now we may re-apply the basic reduction step.
1200: Continuing in this way we get strict decrease of $size(\bv)$ at each
1201: step, with the only possible halting step being the smallest curvature 
1202: equals $0$. Since the size of the curvature vector 
1203: is bounded below and decreases by at least one at each
1204: step, the procedure terminates at $(0,0,g,g)$, up to a permutation.~~~$\bsq$
1205: 
1206: 
1207: \paragraph{Proof of Theorem~\ref{th41n}.}
1208: For existence, the  super-packing generated by
1209: a Descartes configuration with curvature vector $(0,0, g,g)$,
1210: which is a homothetically scaled version of the configuration
1211: $(0,0,1,1)$ pictured in Figure 1, is necessarily integral 
1212: with divisor $g$.
1213: 
1214: For uniqueness,
1215: Theorem~\ref{reduction} shows that any two geometric integral Apollonian
1216: super-packings with divisor $g$ each contain a Descartes configuration
1217: whose curvatures are $(0,0,g,g)$ up to permutation and orientation. 
1218: Now  it is true for any two such Descartes configurations with identical
1219: curvature vectors are congruent,
1220: i.e. one is obtainable from the other by a Euclidean motion.
1221: This is obvious by inspection for the $(0,0,g,g)$ Descartes
1222: configuration, which necessarily
1223: consists  of two touching circles of radius $\frac{1}{g}$
1224: and two parallel lines.
1225: 
1226: Now the Euclidean motion that takes one
1227: Descartes configuration to the other, also takes the super-packing
1228: generated by the first configuration  to
1229: the one generated by the  other, because the super-packing is defined by the
1230: action of the super-Apollonian group on the left on the
1231: Descartes configuration $\bW_{\sD}$, and this commutes with
1232: the Euclidean motion acting as a M\"{o}bius transformation
1233: on the right. This establishes uniqueness.
1234: ~~~$\bsq$
1235: 
1236: We can use the freedom of
1237: a Euclidean motion allowed in Theorem~\ref{th41n}
1238: to make an internal Apollonian super-packing strongly integral.
1239: 
1240: \begin{theorem}~\label{th43n}
1241: For each integral geometric Apollonian super-packing 
1242: there is a Euclidean
1243: motion that takes it to a strongly integral geometric
1244: Apollonian super-packing.
1245: \end{theorem}
1246: 
1247: \paragraph{Proof.}
1248: Using Theorem~\ref{reduction}  
1249: each integral geometric Apollonian packing contains 
1250: a Descartes configuration $\sD$ with curvatures $(0,0, g,g)$; 
1251: note that for a  geometric packing the order and orientation of the
1252: Descartes configuration do not matter.  The curvature vector
1253: determines the Descartes configuration up to 
1254: congruence. We can now find a strongly integral
1255: Descartes configuration $\sD'$ with this curvature vector.
1256: For  $g=1$ such a configuration is given explicitly by
1257: \eqn{N601} below, and for larger $g$ we obtain a strongly
1258: integral configuration from it using the homothety 
1259: $(x, y) \mapsto \frac{1}{g}(x, y)$. There exists a Euclidean
1260: motion that maps $\sD$ to $\sD'$, since they are congruent
1261: configurations. This motion maps the super-packing $\sA^{S}[\sD]$
1262: to the super-packing $\sA^{S}[\sD']$,  which is strongly integral.
1263: ~~~$\bsq$
1264: 
1265: %*****************************************************************************
1266: %
1267: % SECTION 5. 
1268: %
1269: %*****************************************************************************
1270: \section{Strongly Integral Super-packings} 
1271: \setcounter{equation}{0}
1272: 
1273: A Descartes configuration $\sD$ is {\em strongly integral}
1274: if its associated $4 \times 3$
1275: curvature center-coordinate matrix 
1276: $\bM_{\sD}$ is an integer matrix; this property is
1277: independent of ordering or orientation of the
1278: Descartes configuration.
1279: A super-packing is called {\em strongly integral} if
1280: it contains one (and hence all) Descartes configurations 
1281: having 
1282: this property. Since a strongly integral super-packing
1283: is integral, it has a divisor $g$ as an invariant.
1284: 
1285: Our main object in this section is to classify strongly
1286: integral super-packings, as follows.
1287: 
1288: %
1289: %
1290: % theorem 5.1
1291: %
1292: %
1293: \begin{theorem}~\label{orbits}
1294: (1) For each $g \ge 1$ there  are exactly $8$ 
1295: different geometric Apollonian
1296: super-packings that are strongly integral
1297: and have divisor $g$.
1298: 
1299: (2) The set of all ordered, oriented Descartes configurations
1300: that are strongly integral and have a given divisor $g$
1301: fill exactly 384 orbits of the super-Apollonian group.
1302: \end{theorem}
1303: 
1304: This theorem classifies  these super-packings as rigid objects, not
1305: allowed to be moved by Euclidean motions. 
1306: To prove this result we derive a normal form for
1307: a ``super-root quadruple'' in a super-packing of the kind
1308: above, as follows. 
1309: 
1310: %
1311: % theorem 5.2
1312: %
1313: %
1314: \begin{theorem}\label{normal_form} 
1315: Given a strongly integral Apollonian super-packing $\sA^s[\sD_0]$ with the 
1316:  divisor  $g \geq 1$, 
1317: there exists a unique ``reduced'' Descartes configuration 
1318: $\sD \in \sA^S[\sD_0]$ 
1319: whose curvature-center coordinate matrix $\bM=\bM_\sD$ is of the form
1320: $\bA_{m,n}[g]$ or $\bB_{m,n}[g]$ for  $m, n \in \{0, 1\}$,  
1321: up to a permutation of rows,  where
1322: \beq \label{ABmn} 
1323: \bA_{m,n}[g]=\pm \left[ \begin{array}{ccc} 
1324:      0 & 0 & 1 \\
1325:      0 & 0 & -1 \\
1326:      g & m & n \\
1327:      g & m-2 & n \end{array}  \right]
1328: \qquad  
1329: \bB_{m,n}[g]=\pm \left[ \begin{array}{ccc} 
1330:      0 & 1 & 0 \\
1331:      0 & -1 & 0 \\
1332:      g & m & n \\
1333:      g & m & n-2 \end{array}  \right],
1334: \eeq
1335: and the sign is determined by the orientation of $\sD_0$. 
1336: \end{theorem}
1337: \pf For a strongly integral Apollonian super-packing $\sA^S[\sD_0]$ with the 
1338: divisor  $g$, 
1339: by Theorem \ref{reduction} there exists a
1340: strongly integral Descartes configuration $\sD \in \sA_S[\sD_0]$ with
1341: curvatures $\pm(0,0,g,g)$. The two 
1342: straight lines in $\sD$ must be parallel to either $x$-axis or $y$-axis.
1343: It follows that the $4\times 3$ curvature-center coordinate matrix $\bM_\sD$
1344: is of the form $\bA_{m,n}[g]$, or $\bB_{m,n}[g]$, for some $m, n \in \ZZ$, 
1345: up to a permutation of rows. 
1346: 
1347: We now reduce $m, n$ to take the values $0, 1$ using the
1348: following identities, which are easy to check.
1349: \begin{eqnarray*}
1350: \bP_{(34)}\bS_3 \bA_{m,n}[g]=\bA_{m-2, n}[g],  & \qquad&
1351: \bP_{(34)}\bS_3 \bB_{m,n}[g]=\bB_{m, n-2}[g], \\
1352: \bP_{(34)}\bS_4 \bA_{m,n}[g]=\bA_{m+2, n}[g], & \qquad &
1353: \bP_{(34)}\bS_4 \bB_{m,n}[g]=\bB_{m, n+2}[g], \\
1354: \bP_{(12)}\bS_1^T \bA_{m,n}[g]=\bA_{m,n+2}[g], & \qquad&
1355: \bP_{(12)}\bS_1^T \bB_{m,n}[g]=\bB_{m+2,n}[g], \\
1356: \bP_{(12)}\bS_2^T \bA_{m, n}[g]=\bA_{m, n-2}[g],  & \qquad &
1357: \bP_{(12)}\bS_2^T \bB_{m, n}[g]=\bB_{m-2, n}[g].
1358: \end{eqnarray*}
1359: where $\bP_{(ij)}$ is the permutation matrix that exchanges $i$ and $j$. 
1360: Also note that  $\bP_\sigma \bS_i=\bS_{\sigma(i)} \bP_\sigma$,
1361: $\bP_\sigma \bS_i^T =\bS_{\sigma(i)}^T \bP_\sigma$.  Hence there is
1362: a series of group operations in $\sA^S$  which takes $\bM_\sD$ to 
1363: a permutation of $\bA_{m,n}[g]$ or $\bB_{m,n}[g]$ with $m, n \in \{0, 1\}$. 
1364: This proves the existence of the ``reduced'' Descartes configuration 
1365: in the Apollonian super-packing $\sA^S[\sD_0]$. 
1366: 
1367: To prove the uniqueness, it suffices to show that the 
1368: $24 \times 8 \times 2=384$ 
1369: Descartes configurations whose curvature-center coordinate matrices
1370: are 
1371: $$
1372: \{ \bP \bA_{m,n}[g], \bP \bB_{m,n}[g] \ | \ \bP \in \text{Perm}_4, m, n \in \{0, 1\}
1373: \}
1374: $$ 
1375: are in different Apollonian super-packings. 
1376: (There are two signs for each of $\bA_{m,n}[g]$ and $\bB_{m,n}[g]$. )
1377: In what follows we let $\tilde{\bA}_{m,n}[g]$, and $\tilde{\bB}_{m,n}[g]$
1378: denote the unique $4 \times 4$ augmented curvature-center coordinate
1379: matrices extending $\bA_{m,n}[g]$ and $\bB_{m,n}[g]$, respectively;
1380: uniqueness holds  by Theorem 3.1 of part I.
1381: 
1382: Note that each $\bS_i$ and $\bS^T_i$ preserves the (total) orientation 
1383: of the Descartes configuration, as well as the parity of every 
1384: element of $\bM_\sD$. 
1385: First we show that the matrices 
1386: $\bA_{m,n}[g], \bB_{m,n}[g], (m,n \in \{0,1\}$ 
1387: are in distinct orbits of $\sA^S \times \text{Perm}_4$.
1388: To see this, for each integral vector $\bv \in \zz^4$, 
1389: let $\kappa(\bv)$ be the
1390: number of even terms in $\bv$, and
1391: for any strongly integral Descartes configuration $\sD$ 
1392: let 
1393: $$
1394: \kappa(\bM_\sD)=(\kappa(\bv_1), \kappa(\bv_2), \kappa(\bv_3)),
1395: $$
1396: where 
1397: $\bv_1, \bv_2, \bv_3$ are the column vectors of $\bM_\sD$.
1398: Then $\kappa(\bM_\sD)$ is invariant under the action of 
1399: $\sA^S \times \text{Perm}_4$.  For $m, n \in \{0, 1 \}$,
1400: $\kappa(\bA_{m,n}[g]), \kappa(\bB_{m,n}[g])$ are all distinct except
1401: $\kappa(\bA_{1, 0}[g])=\kappa(\bB_{0, 1}[g])=(*, 2, 2)$, where
1402: $*$ is $2$ if $g$ is odd, and $4$ if $g$ is even. However, 
1403: $\bA_{1,0}[g]$ and $\bB_{0,1}[g]$ can not be equivalent under the action of 
1404: $\sA^S \times \text{Perm}_4$. Arguing by contradiction, 
1405: assume that there exists a
1406: matrix $\bU \in \sA^S \times \text{Perm}_4$ such that
1407: $\bU \bA_{1, 0}[g]=\bB_{0, 1}[g]$.  This relation lifts
1408: to augmented curvature-center coordinates:  
1409: $\bU \tilde{\bA}_{1,0}[g]=\tilde{\bB}_{0,1}[g]$. 
1410: It follows that 
1411: $\bU=(\tilde \bA_{1,0}[g])^{-1} \tilde \bB_{0,1}[g]$ is unique.
1412: We can directly verify that for $m,n \in \ZZ$ 
1413: 
1414: \beq \label{aug-ABmn} 
1415: \tilde \bA_{m,n}[g]=\left[\begin{array}{cccc}
1416: 2(n+1)/g &  0 & 0 &  1  \\
1417: 2(1-n)/g &  0 & 0 & -1 \\
1418: (m^2+n^2-1)/g &   g & m & n \\
1419: ((m-2)^2+n^2-1)/g &  g & m-2 & n \end{array}\right],
1420: \nonumber \\
1421: \tilde \bB_{m,n}[g]=\left[\begin{array}{cccc}
1422: 2(m+1)/g &   0 & 1 & 0 \\
1423: 2(1-m)/g &   0 & -1 & 0 \\
1424: (m^2+n^2-1)/g &   g & m & n \\
1425: (m^2+(n-2)^2-1)/g  &   g & m & n-2 \end{array}\right], 
1426: \eeq
1427: by checking that these satisfy the identity of Theorem 3.2 of
1428: part I necessary and sufficient 
1429: to be augmented curvature-center coordinates.
1430: Now it is easy to verify that 
1431:  $\bP_{(14)}\bP_{(23)}\bD \tilde \bA_{1, 0}[g]=\tilde \bB_{0, 1}[g]$, 
1432: where $\bD=-\bQ_D$ is defined as in \S3 of Part I \cite{GLMWY11}, 
1433: \beq \label{matrixD}
1434: \bD=\frac{1}{2} \left[ \begin{array}{rrrr} 
1435:  -1 & 1 & 1 & 1  \\
1436:  1 & -1 & 1 & 1  \\
1437:  1 & 1 & -1 & 1 \\
1438:  1 & 1 & 1 & -1 \end{array} \right].
1439: \eeq 
1440: By the uniqueness 
1441: $\bP_{(14)}\bP_{(23)} \bD =\bU \in \sA^S \times \text{Perm}_4$, 
1442: which is  impossible since $\sA^S \times \text{Perm}_4$ consists of 
1443: integral matrices only, while $\bD$ has half integers.
1444: In conclusion, $\bA_{m, n}[g], \bB_{m, n}[g]$,
1445: ($m, n \in  \{0, 1\}$) are in distinct orbits of
1446: $\sA^S \times \text{Perm}_4$. 
1447: 
1448: The final step
1449: is to show that for any permutation $\bP \neq \bI$, $\bP \bA_{m,n}[g]$
1450: (resp. $\bP \bB_{m,n}[g]$) 
1451: can not be obtained from $\bA_{m,n}[g]$ (resp. $\bB_{m,n}[g]$) 
1452: by an action of $\sA^S$. That is, we claim: 
1453:  if for  a permutation matrix $\bP \in \text{Perm}_4$,
1454: there exists a matrix $\bU \in \sA^S$ such that
1455: \[
1456: \bU \bA_{m,n}[g] =\bP \bA_{m,n}[g], \qquad \text{or } \bU \bB_{m,n}[g] =\bP \bB_{m,n}[g],
1457: \]
1458: then $\bP=\bI$.
1459: 
1460: To establish the claim, 
1461: consider  again the $4 \times 4$ ACC-coordinate matrices $\tilde \bA_{m,n}[g]$ 
1462: and $\tilde \bB_{m,n}[g]$. From \S3.1 of Part I \cite{GLMWY11},  
1463: for any Descartes configuration $\sD$, 
1464: the curvature-center coordinate matrix $\bM_\sD$  
1465: can be uniquely extended to a $4 \times 4$ ACC-coordinate matrix
1466: $\bW_\sD$. It follows that  
1467: if $\bU \bA_{m,n}[g] =\bP \bA_{m,n}[g]$, then
1468: the equality holds for their $4\times 4$ ACC-coordinate matrices, i.e., 
1469: $\bU \tilde \bA_{m,n}[g] =\bP \tilde \bA_{m,n}[g]$.  It implies  
1470:  $\bU=\bP \in \sA^S \cap \text{Perm}_4$. 
1471: However, comparing the size of $\bU$ and $\bP$, where the size of 
1472: a matrix $\bU$ is defined as $f(\bU):= \bo^T \bU \bo$, we have
1473: $f(\bP):={\bf 1}^T \bP {\bf 1}=4$ for $\bP \in \text{Perm}_4$,
1474: and $f(\bU):={\bf 1}^T \bU {\bf 1} \geq 8$ for any $\bU \in \sA^S, 
1475: \bU\neq \bI$,
1476: (c.f.\S5 of Part I \cite{GLMWY11}).  Therefore 
1477: the only possibility is $\bU=\bP=\bI$. The same argument applies to 
1478: $\bB_{m,n}[g]$, and the claim follows. 
1479: 
1480: We conclude that a reduced Descartes
1481: configuration of the
1482: form \eqn{aug-ABmn} in any 
1483: strongly integral Apollonian super-packing 
1484: exists and is unique. ~~~\bsq 
1485: 
1486: 
1487: 
1488: \paragraph{Proof of Theorem~\ref{orbits}.}
1489: (1) Since there are 48 orbits of ordered, oriented
1490: Descartes configurations corresponding to each
1491: geometric super-packing, to show there are exactly
1492: $8$ geometric super-packings it suffices to  show 
1493: that the strongly integral Descartes configurations
1494: form  $384$ orbits of the Apollonian group,
1495: which we do below.
1496: 
1497:  
1498: (2) We enumerate  the complete set of 
1499: ordered, oriented Descartes configurations that
1500: are strongly integral, with greatest common divisor $g$,
1501: as follows. 
1502: By Theorem \ref{normal_form},
1503: any such Descartes configuration  
1504: is equivalent under the action of $\sA^S$ to a permutation of a Descartes 
1505: configuration whose $4\times 3$ curvature-center coordinate matrix is of 
1506: the form $A_{m,n}$ or $B_{m,n}$, with $m,n \in \{0, 1\}$. 
1507: The uniqueness of Theorem \ref{normal_form} asserts that the $24$ permutations
1508: of $A_{m,n}[g]$ ($B_{m,n}[g]$) are all in distinct orbits of the 
1509: super-Apollonian
1510: group. Considering the two choices of orientations, we get 
1511: $24 \times 8 \times 2=384$ orbits. $~~~\bsq$ \\
1512: 
1513: 
1514: 
1515: %*****************************************************************************
1516: %
1517: % SECTION 6
1518: %
1519: %****************************************************************************
1520: 
1521: \section{Primitive Strongly Integral Super-packings}
1522: \setcounter{equation}{0}
1523: 
1524: The strongly integral
1525: superpackings having  a given curvature g.c.d. $g$ are each obtainable
1526: from one  with $g=1$ by homothety.
1527: A homothety $(x,y) \mapsto  r(x,y)$ changes all curvatures
1528: by $\frac{1}{r}$ while leaving (curvature)$\times$(center)
1529: unchanged. Applying the homothety with $r = \frac{1}{g}$ 
1530: takes a strongly integral super-packing with g.c.d. $g$ to
1531: one with g.c.d. equal to $1$.
1532: 
1533: We now  consider strongly integral super-packings
1534: having $g=1$, which we call {\em primitive} super-packings.
1535: Results for them carry over easily to those with divisor $g> 1$ by 
1536: applying a homothety. Theorem~\ref{orbits} showed there are
1537: exactly $8$ such packings. 
1538: 
1539: For convenience we will single
1540: out a particular one of them and term it the {\em standard strongly
1541: integral super-packing}.
1542: We choose this to be the super-packing generated
1543: by the ordered, oriented Descartes configuration
1544: having 
1545: \begin{equation}~\label{N601}
1546: \bM_{\sD_1}= \left[
1547: \begin{array}{rrr}
1548:  0 & 0 & 1 \\
1549:  0 & 0 & -1 \\
1550:  1 & 1 & 0 \\
1551:  1 & -1 & 0
1552: \end{array}
1553: \right]~~~~~~\mbox{so~that}~~~~~~~~
1554: \bW_{\sD_1}= \left[
1555: \begin{array}{rrrr}
1556: 2 &  0 & 0 & 1 \\
1557: 2 &  0 & 0 & -1 \\
1558: 0 &  1 & 1 & 0 \\
1559: 0 &  1 & -1 & 0
1560: \end{array}
1561: \right].
1562: \end{equation}
1563: This corresponds to a $(0,0,1,1)$ Descartes
1564: quadruple, with the centers of the two circles
1565: lying along the $x$-axis and the circles touching
1566: at the origin $(0,0)$. The associated geometric
1567: integral super-packing is the one pictured in
1568: \S3. Results proved below for
1569: the standard  super-packing apply generally to all eight 
1570: primitive integral super-packings, using the Euclidean
1571: motions mapping between them described after the
1572: proof of Theorem~\ref{th61} below. 
1573: 
1574: We first show  that the geometric standard 
1575: strongly-integral super-packing
1576: has a large group of symmetries, which form  a 
1577: crystallographic group of the plane.
1578: %
1579: %
1580: % Theorem 6.1
1581: %
1582: %
1583: 
1584: \begin{theorem}~\label{th61}
1585: The geometric standard strongly integral super-packing is
1586: invariant under the following Euclidean motions:
1587: 
1588: (1) The lattice of translations $(x,y) \mapsto (x+2, y)$, 
1589: and $(x,y)  \mapsto (x, y+2)$,
1590: 
1591: (2) The reflections $(x,y) \mapsto (-x, y)$, and 
1592: $(x,y)  \mapsto (x, -y)$.
1593: 
1594: \noindent The crystallographic group generated by
1595: these motions is  the complete set of
1596: Euclidean motions leaving the geometric standard
1597: strongly integral super-packing invariant.
1598: \end{theorem}
1599: 
1600: \paragraph{Proof.}
1601: The key fact used is that the action of the super-Apollonian
1602: group on Descartes configurations commutes with the
1603: action of Euclidean motions acting on Descartes
1604: configurations as M\"{o}bius transformations. 
1605: This was shown in part I \cite[Theorem 3.3(4)]{GLMWY11}.
1606: 
1607: (1) There is a 
1608: Descartes configuration 
1609: corresponding to $\sD_0$ shifted by $2$ in the 
1610: $x$-direction and the $y$-direction; call them $\sD_0^{x}$
1611: and $\sD_0^{y}$, respectively. These are given by the
1612: actions of $\bS_{4}$ and  $\bS_{1}^{\perp}$, respectively.
1613: Treating the $x$-shift first, we then have 
1614: $$
1615: \sA^{S}[\sD_0] = \sA^{S}[\sD_0^{x}]= \sA^{S}[ \ft_{x}(\sD_0^{'})],
1616: $$
1617: in which  $\ft_{x}: \bz \mapsto \bz +2$ is 
1618: the Euclidean motion translation by $\bv=(2,0)$
1619: as in  Appendix A of part I, and the
1620: ordered, oriented Descartes configuration $\sD_0^{'}$ is
1621: a permutation of $\sD_0$. 
1622: Then the  geometric super-packing
1623: associated to $\sA^{S}[\sD_0^{'}]$
1624: is therefore identical with that of $\sA^{S}[\sD_0]$, and that of 
1625: $\sA^{S}[ \ft_{x}(\sD_0^{'})]$ translates it by $(2,0)$.
1626: Thus the geometric packing is invariant under this translation.
1627: The argument for translation by $(0,2)$ is similar. 
1628: 
1629: (2) This geometric Descartes  configuration $\sD_0$ 
1630: is invariant under the reflections $(x,y) \mapsto (-x, y)$, and 
1631: $(x,y)  \mapsto (x, -y)$ viewed as M\"{o}bius transformations.
1632: The effect of these transformations on the ordered,
1633: oriented Descartes configuration is to permute its rows.
1634: It follows as in (1) that the associated geometric 
1635: Apollonian super-packings  are identical.
1636: 
1637: To see that these motions generate the full group of Euclidean
1638: motions leaving the super-packing invariant, we observe
1639: that the full group acts discontinuously on the plane. 
1640: This is because the image of the $(0,0,1,1)$ configuration
1641: is either left fixed, or else it moves a distance of at least two
1642: in some direction, so that its circles do not overlap. 
1643: Thus it must be contained in a crystallographic group whose
1644: translation subgroup is given by (1) above. Now the only
1645: possibilities are to extend the group by a subgroup of
1646: the finite point group of motions leaving $(0,0)$ fixed (of order $8$) 
1647: leaving
1648: the lattice $\ZZ[(0,2), (2,0)]$ of translations invariant. 
1649: Here (2) gives an extension of order $4$. No larger
1650: extension occurs by observing that otherwise the image
1651: of the $(0,0, 1,1)$ and $(-1, 2, 2, 3)$ configurations at
1652: the origin would cross themselves.
1653: $~~~\bsq$ \\
1654: 
1655: One can now check that 
1656: the  $8$ primitive geometric strongly integral super-packings
1657: given by Theorem~\ref{orbits} are obtained from
1658: the standard strongly integral
1659: super-packing by $8$ cosets of the Euclidean motions 
1660: $(x, y) \mapsto (x+1, y)$,
1661: $(x,y) \mapsto (x, y+1)$ and $(x,y) \mapsto (y, x)$
1662: with respect to the symmetries in Theorem~\ref{th61}.
1663: They are specified by the location 
1664: and orientation of the $(0,0,1,1)$
1665: configuration.
1666: 
1667: Our next result  shows that every primitive integral
1668: Descartes configuration with no curvature zero
1669: occurs inside the geometric standard strongly integral super-packing
1670: in a specified location.
1671: 
1672: %
1673: % Theorem 6.2
1674: %
1675: %
1676:  
1677: \begin{theorem}~\label{th62}
1678: In the geometric standard strongly integral super-packing,
1679: for each (unordered) primitive integral Descartes quadruple
1680: $(a, b, c, d)$ except for $(0,0,1,1)$, 
1681: there exists a Descartes configuration
1682: having these curvatures, such that the
1683: center of its largest circle lies in
1684: the closed unit square
1685: $\{(x,y): 0 \le x \le 1,~ 0 \le y \le 1\}$.
1686: The location of the center of this largest circle is unique.
1687: \end{theorem}
1688: 
1689: \paragraph{Proof.}
1690: To establish  existence,
1691: we first  show that a Descartes configuration
1692: of the curvatures occurs somewhere inside 
1693: the standard integral super-packing. This holds
1694: because the super-packing generated by such
1695: a configuration is an integral super-packing
1696: with divisor $1$, which by Theorem~\ref{th41n} 
1697: is unique up to a Euclidean motion. Thus the
1698: standard strongly integral super-packing must
1699: contain an isometric copy of it. Now that
1700: we have such a configuration inside the packing,
1701: we can use  
1702: the translation symmetries in Theorem~\ref{th61}
1703: to move it so that its largest circle
1704: has center 
1705: inside the half-open square $\{ (x, y) : -1 \le x < 1, -1 \le y < 1\}$.
1706: If we have $-1\le x < 0$ then we apply the symmetry
1707: $(x,y) \mapsto (-x,y)$, while if $-1\le y < 0$ we
1708: apply $(x,y) \mapsto (x,-y)$, as necessary.
1709: 
1710: To establish uniqueness, we argue by contradiction.
1711: If uniqueness failed,  there would exist a Euclidean motion taking
1712: one of these Descartes configurations to the other. 
1713: Since a single Descartes configuration generates the
1714: entire super-packing, we conclude that the super-packing
1715: is left invariant under this extra Euclidean motion. 
1716: Theorem~\ref{th61} described all such automorphisms
1717: and all of them, except the identity,  
1718: map every point in the interior of the unit square 
1719: square strictly outside the square. This 
1720: contradicts the assumption that the center of
1721: the largest circle of the first configuration
1722: is mapped to that of the second, when
1723: at least one of these points is strictly  inside the
1724: unit square. In the remaining cases where both 
1725: centers lie on the boundary, one shows 
1726: they are must lie  the same
1727: boundary edge and that the automorphism leaves this
1728: edge fixed, so they are identical. 
1729: Note that this  argument shows in passing
1730: that the largest circle is unique, once $(0,0,1,1)$
1731: is excluded. 
1732: $~~~\bsq$ \\
1733: 
1734: We come now to the main result of this section,
1735: which asserts that the 
1736: geometric standard super-packing
1737: contains a copy of every integral Apollonian circle
1738: packing in a canonical way. The circles in the geometric
1739: standard super-packing can be foliated into 
1740: a union of geometric Apollonian packings.  
1741: 
1742: %
1743: % Theorem 6.3
1744: %
1745: %
1746: 
1747: \begin{theorem}~\label{th63}
1748: (1) Each circle in the standard super-packing
1749: with center inside the 
1750: half-open unit square
1751: $\{(x,y): 0 \le x < 1,~ 0 \le y < 1\}$
1752: is the exterior boundary circle of a unique 
1753: primitive integral Apollonian circle packing contained
1754: in the geometric standard integral super-packing.
1755: 
1756: (2) Every primitive integral Apollonian circle packing,
1757: except for the packing  $(0,0,1,1)$,
1758: occurs exactly once in this list. 
1759: \end{theorem}
1760: 
1761: \paragraph{Proof.}
1762: (1) Let the circle $C$ be given.
1763: Recall from the proof of Theorem~\ref{th31}
1764: that  there is a unique minimal length admissible sequence
1765: $\bU_m \bU_{m-1} \cdots \bU_1 [\sD_0]$  
1766: of generators of the super-Apollonian
1767: group that yield a Descartes configuration $\sD$ containing
1768: the given circle $C$, say in its $j$-th position.
1769: Admissibility
1770: requires that  $\bU_m = \bS_j$ or $\bS_i^{\perp}$ for some $i \ne j$.
1771: Then multiplying by $\bS_j^{\perp}$ also gives an 
1772: admissible sequence, and the Descartes configuration
1773: $$
1774: \sD':=\bS_j^{\perp}[\sD] =  \bS_j^{\perp} \bU_m \bU_{m-1} \cdots \bU_1 [\sD_0]
1775: $$
1776: consists of the circle
1777: $C$  plus three new circles nested inside the interior of $C$.
1778: This Descartes configuration $\sD'$ generates 
1779: an Apollonian packing having the circle $C$ as outer
1780: boundary, contained in the standard strongly integral super-packing. It
1781: is unique, because if there were a second Apollonian packing
1782: inside the bounding circle it would contain circles crossing
1783: those in the first packing, contradicting Theorem~\ref{th31}.
1784: 
1785: (2) Recall from \cite[Sect. 3 and 4]{GLMWY2}
1786: that inside  each integer Apollonian packing is a 
1787: positively oriented Descartes configuration
1788: whose absolute values of curvatures $(a, b, c, d)$ are minimal, which is
1789: called a {\em root quadruple}. 
1790: Theorem 4.1 of \cite{GLMWY2} showed that aside   from the root quadruple
1791: $(0,0,1,1)$ every root quadruple is of the form 
1792: $a< 0< b \le c \le d$. Root quadruples are characterized
1793: by satisfying the extra condition
1794: \beql{603root}
1795: a+ b+c  \ge d.
1796: \eeq
1797: All the circles in
1798: the resulting Apollonian packing are contained inside
1799: a {\em bounding circle} of curvature $N=|a|$, i.e., radius $\frac{1}{N}$.
1800: Theorem~\ref{th62} shows that for each root quadruple,
1801: with all curvatures nonzero 
1802: there is a matching Descartes configuration
1803: whose largest circle has center inside the unit square
1804: and this largest circle is unique. The Apollonian
1805: packing contained inside this circle by  is the integral packing with
1806: the given root quadruple, and it is unique
1807: by the result of (1).  (In some cases, like $(-1, 2,2, 3)$
1808: the root configuration is not unique, but the root quadruple
1809: and the packing itself
1810: are  always unique.) Thus every primitive integer Apollonian
1811: circle packing, except $(0,0,1,1)$, occurs exactly once bounding
1812: circles having center in  the closed unit square.
1813: $~~~\bsq$ \\
1814: 
1815: 
1816: 
1817: 
1818: 
1819: The initial part of the standard  super-packing to depth 200 
1820: inside the unit square is pictured in Figure~\ref{fig8} below.
1821: 
1822: 
1823: %
1824: % Figure 8  (superpacking to depth 200)
1825: %
1826: %
1827: \begin{figure}[htbp]
1828: \centerline{\epsfxsize=6.0in \epsfbox{0mod1.ps}}
1829: \caption{A ``deeper'' initial  part of a super-packing
1830: (square of sidelength  $1$)}~\label{fig8}
1831: \end{figure}
1832: 
1833: 
1834: One can make  further 
1835: computer experiments plotting the circles having various
1836: curvatures restricted $(\bmod~4)$ inside the unit
1837: square.  The results for circles
1838: having curvatures $1~(\bmod~2)$, 
1839: $2~(\bmod~4)$  and $0~(\bmod~4)$ and size at most 
1840: $200$ are pictured in the following three figures.
1841: 
1842: %
1843: %  Figure 9: (1 mod 2)
1844: %
1845: %
1846: 
1847: \begin{figure}[htbp]
1848: \centerline{\epsfxsize=6.0in \epsfbox{1mod2.ps}}
1849: \caption{Circles of curvature $1~(\bmod~2)$}~\label{fig9}
1850: \end{figure}
1851: 
1852: %
1853: %  Figure 10 : (2 mod 4)
1854: %
1855: %
1856: 
1857: \begin{figure}[htbp]
1858: \centerline{\epsfxsize=6.0in \epsfbox{2mod4.ps}}
1859: \caption{Circles of curvature $2~(\bmod~4)$}~\label{fig10}
1860: \end{figure}
1861: 
1862: %
1863: %  Figure 11 : (0 mod 4)
1864: %
1865: %
1866: 
1867: \begin{figure}[htbp]
1868: \centerline{\epsfxsize=6.0in \epsfbox{0mod4.ps}}
1869: \caption{Circles of curvature $0~(\bmod~4)$}~\label{fig11}
1870: \end{figure}
1871: 
1872: The figures empirically
1873: indicate that the following extra reflection symmetries occur;
1874: the line of symmetry is indicated on the figure  with
1875: a dotted line.
1876: 
1877: (a) $1~(\bmod~2)$ circles are symmetrical under reflection
1878: in the line $x= 1-y$.
1879: 
1880: (b) $2~(\bmod~4)$ circles are symmetrical under reflection
1881: in the horizontal line $y= \frac{1}{2}$.
1882: 
1883: (c) $0~(\bmod~4)$ circles are symmetrical under reflection
1884: in the vertical line $x = \frac{1}{2}$.
1885: 
1886: 
1887: Based on this experimental evidence,
1888: one  of the authors (CLM) conjectured that
1889: these symmetries hold.
1890: They  were subsequently proved by
1891: S. Northshield ~\cite{No03}.
1892: 
1893: We may also illustrate these symmetry properties at the
1894: level of circles of a fixed curvature. Recall from \S3
1895: that each such circle contains a unique Apollonian packing
1896: having it as outer circle. Such a circle
1897: can then be  labelled by  root quadruple in the
1898: sense of \cite{GLMWY2} of this integral Apollonian packing.  
1899: The cases of curvature $-6, -8,$ and $-9$ corresponding to
1900: the three cases above were pictured in \S2.
1901: In our paper considering number-theoretic
1902: properties of integral Apollonian packings, it was shown
1903: in  \cite[Theorem 4.2]{GLMWY2} that the number of distinct primitive
1904: integral Apollonian packings with a given curvature $-n$ 
1905: of the outer circle had an interpretation as a class number
1906: $h^{\pm}(-4n^2)$ of positive definite binary quadratic forms of discriminant
1907: $\Delta= -4n^2$, under $GL(2, \zz)$-equivalence. One can raise 
1908: the  question whether there is some interpretation
1909: of the extra symmetries (a)-(c) above in terms of the 
1910: associated class group structure under $SL(2, \zz)$ equivalence. 
1911: 
1912: 
1913: 
1914: 
1915: 
1916: 
1917: 
1918: %*****************************************************************************
1919: %
1920: % SECTION 7
1921: %
1922: %*****************************************************************************
1923: %
1924: \section{The Super-Apollonian Group has Finite Covolume}
1925: \setcounter{equation}{0}
1926: In this section we show that the super-Apollonian group
1927: is of finite volume as a discrete subgroup of
1928: the real Lie group 
1929: $Aut(Q_{\sD}) \simeq O(3, 1)$. This 
1930: follows from  the fact that
1931: the integer Lorentz group $O(3, 1; \ZZ)$ has
1932: finite covolume in $O(3, 1)$, and the following result.
1933: 
1934: %
1935: %
1936: % theorem 7.1
1937: %
1938: %
1939: 
1940: \begin{theorem}~\label{th51} 
1941: (1) The super-Apollonian group $\sA^{S}$ is a normal
1942: subgroup of index 48 in the group $Aut(Q_{D}, \ZZ)$.
1943: The group $Aut(Q_{D}, \ZZ)$ is generated by
1944: the super-Apollonian group and the finite
1945: group of order 48 generated by the $4 \times 4$
1946: permutation matrices and $\pm \bI$.
1947: 
1948: (2) The super-Apollonian group $\sA^{S}$is a normal subgroup
1949: of index 96 in the group
1950: $G= \bJ_0~ O(3, 1; \ZZ) \bJ_0^{-1}$, where
1951: $$
1952: \bJ_0=\frac{1}{2} \left[ \begin{array}{rrrr}
1953: 1 & 1 & 1 & 1 \\
1954: 1 & 1 & -1 & -1 \\
1955: 1 & -1 & 1 & -1 \\
1956: 1 & -1 & -1 & 1 \end{array} \right].
1957: $$
1958: The group $G$ is generated by $Aut(Q_{D}, \ZZ)$
1959: and the duality matrix $\bD$, and consists
1960: of matrices with integer and half-integer entries.
1961: \end{theorem}
1962: 
1963: The duality matrix $\bD$ is given by (\ref{matrixD}) in \S5.
1964: This result  allows us to associate to the
1965: super-Apollonian group the normal
1966: subgroup 
1967: $\tilde{\sA}^{S} := \bJ_0^{-1} \sA^{S} \bJ_0$
1968: of the integer Lorentz group $O(3, 1; \ZZ)$, of index $96$.
1969: 
1970: The proof of Theorem \ref{th51} will be
1971: derived in a series of four lemmas.  
1972: Let $\Gamma$ be the group generated by adjoining to $\sA$
1973: the elements of the finite group of order $48$ given by
1974: $\text{Perm}_4$ and $\pm I$, and let
1975: $\tilde \Gamma=< \Gamma, \bD>$. 
1976: The lemmas will  prove that $\Gamma= Aut(Q_D, \zz)$
1977: and $\tilde \Gamma=\bJ_0~ O(3, 1; \ZZ) \bJ_0^{-1}$.
1978: Then analysis of cosets of $\sA$ in these groups 
1979: permits determining the index of $\sA$ in these groups
1980: and showing normality. 
1981: 
1982: To determine $\Gamma$ and $\tilde{\Gamma}$ we will show
1983: $$
1984: \Gamma \leq Aut(Q_D, \zz) \leq \tilde{\Gamma} = G.
1985: $$
1986: It is easy to show that  
1987: $\Gamma$ is a subgroup of $\tilde \Gamma$ of index 2,
1988: and $\bD \in \tilde \Gamma$ but $\bD \notin Aut(Q_D, \zz)$,
1989: so we get  
1990: $Aut(Q_D, \zz) =\Gamma =<\sA^S, \text{Perm}_4, \pm \bI>$. 
1991: It is easy to check the
1992: inclusion  $\Gamma \leq Aut(Q_D, \zz)$ since the generators 
1993: $\bS_i, \bS^T_i$, $\bP_\sigma$ and $\pm \bI$ of $\Gamma$ are all in 
1994: $Aut(Q_D, \zz)$. The inclusion $Aut(Q_D, \zz) \leq G$ is proved in
1995: Proposition \ref{lemma52}.  The equation $G=\tilde \Gamma$ is proved 
1996: in the next three lemmas.
1997: In Lemma \ref{lemma53} we prove that the integer Lorentz group is exactly 
1998: the group $Aut(L_\zz)$ of 
1999: invertible linear transformations that leave the integral Lorentz cone 
2000: $L_\zz$ invariant.  Lemma \ref{lemma54} states that in order to show 
2001: that a group $\sG$ of invertible 
2002: linear transformations of $L_\zz$ equals $Aut(L_\zz)$, 
2003: it is enough to show that (1) the action of $G$ on $L_\zz$ is transitive, 
2004: and (2) there exists a point $v \in L_\zz$ such that the stabilizer 
2005: $\sS_v =\{ \bU \in Aut(L_\zz) \ | \ \bU v =v \}$ is a subset of $\sG$. 
2006: Using Lemma \ref{lemma54}, we check that (1) the action of 
2007: $\bJ_0 \tilde \Gamma \bJ$ on $L_\zz$ is transitive, and (2), 
2008: $\bJ_0 \tilde \Gamma \bJ$ contains the stabilizer $\sS_v$ of the point
2009: $v =(1,1,0,0) \in L_\zz$. This proves $G= \tilde \Gamma$. 
2010: 
2011: %
2012: %
2013: % Lemma 7.2
2014: %
2015: 
2016: \begin{lemma} \label{lemma52} It is true that
2017: $$Aut(Q_D, \zz) \leq G= \bJ_0 O(3,1;\zz) \bJ_0^{-1}.$$
2018: \end{lemma}
2019: \pf
2020: Let $\bU \in Aut(Q_D, \zz)$. We need to show that the entries of
2021: $\bJ_0^{-1} \bU \bJ_0 =\bJ_0 \bU \bJ_0$ are all integers. Since
2022: $\bJ_0= \frac{1}{2}\bo \bo^T -\bT$, where
2023: \[
2024: \bT=\left[ \begin{array}{cccc}
2025: 0 & 0 & 0 & 0 \\
2026: 0 & 0 & 1 & 1 \\
2027: 0 & 1 & 0 & 1 \\
2028: 0 & 1 & 1 & 0 \end{array} \right],
2029: \]
2030: we have
2031: \beq
2032: \bJ_0 \bU \bJ_0 = \frac{1}{4}(\sum_{i,j} U_{ij})\bo \bo^T -\frac{1}{2}\bo \bo^T
2033: \bU \bT -\frac{1}{2} \bT \bU \bo \bo^T+\bT \bU \bT,
2034: \eeq
2035: where $\bT \bU \bT$ is an integer matrix.
2036:                                                                                 
2037: From $\bU^T \bQ_D \bU=\bQ_D$ and $ \bQ_D=\frac{1}{2}(2\bI-\bo \bo^T)$, we have
2038: \beq ~\label{6.2-for1}
2039: \bU ^T (2\bI-\bo \bo^T) \bU=2\bI -\bo \bo^T.
2040: \eeq
2041: Denote by $\bv_i$ the $i$-th column of $\bU$,
2042: and by $size(\bv):=\bo^T \bv$ the size of a vector $\bv$.  Equating the entries
2043: of \eqref{6.2-for1} we get
2044: $2 \bv_i \cdot \bv_j -size(\bv_i)size(\bv_j) =2\delta_{ij}-1$.
2045: In particular, $size(\bv_i)$, $i=1,2,3, 4$  are  odd integers. It follows that
2046: the matrix $\frac{1}{2}\bo \bo^T \bU \bT$ is integral.
2047: 
2048: Note that $Aut(Q_D, \zz)$ is closed under transposition. This is because
2049: \beq \label{6.2for-2}
2050: & & \bU \in Aut(Q_D, \zz) \Longrightarrow \bU^T \bQ_D \bU=\bQ_D
2051:  \Longrightarrow \bU^T \bQ_D \bU \bQ_D \bU^T=\bU^T  \nonumber \\
2052: & \Longrightarrow & \bU \bQ_D \bU^T =( \bU^T \bQ_D)^{-1} \bU^T=(\bQ_D)^{-1}=\bQ_D
2053: \Longrightarrow \bU^T \in Aut(Q_D, \zz).
2054: \eeq
2055: Applying the same argument of the preceding paragraph to $\bU^T$, we
2056: then prove that the matrix $\frac{1}{2} \bT \bU \bo \bo^T$ is integral.
2057: 
2058: 
2059: Again using $2 \bv_i \cdot \bv_j -size(\bv_i)size(\bv_j) =2\delta_{ij}-1$, 
2060: summing over
2061: $i, j =1, \dots, 4$, we get
2062: \[
2063: (\sum_{i,j} U_{i,j})^2=2(\bv_1+\bv_2+\bv_3+\bv_4)\cdot 
2064: (\bv_1+\bv_2+\bv_3+\bv_4) +8.
2065: \]
2066: Note that $\bv_1+\bv_2+\bv_3+\bv_4=(size(\br_1), size(\br_2), size(\br_3), 
2067: size(\br_4))^T$
2068: where $\br_i$ is the $i$-th row of the matrix $\bU$. The sum
2069: $\sum_i (size(\br_i))^2$
2070: is a multiple of 4 since each $ size(\br_i)$ is odd. It follows that
2071: $(\sum_{i,j} U_{i,j})^2$ is a multiple of 8. Since $\sum_{i,j} U_{i,j}$
2072: is an integer, we conclude that $\sum_{i,j} U_{i,j}$ is a multiple of 4
2073: and the matrix  $\frac{1}{4}(\sum_{i,j} U_{ij})\bo \bo^T $ is integral.
2074: This proves Lemma  \ref{lemma52}.  ~~\bsq
2075: 
2076: 
2077: Clearly $\sA^S$, $\text{Perm}_4$ and $\pm \bI$ are subgroups
2078: of $Aut(Q_D, \zz)$.  Let $\Gamma$ be the group generated by $\sA^S$,
2079: $\text{Perm}_4$ and $\pm \bI$, and let 
2080: $\tilde \Gamma := \langle\Gamma, \bD \rangle$.
2081: Then $\Gamma$ is a subgroup of $\tilde \Gamma$ of index $2.$
2082:                                                                                 
2083: The {\em Lorentz light cone} is the set of points
2084: $\{ (y_0, y_1, y_2, y_3)^T \in \rr^4 ~: \ -y_0^2+y_1^2+y_2^2+y_3^2 = 0\}$.
2085:  Let $L_\zz$ be the set of integer points in the
2086: Lorentz light cone, i,e,
2087: $$L_\zz :=\{ (y_0, y_1, y_2, y_3)^T \in \zz^4 ~:~ \ -y_0^2+y_1^2+y_2^2+y_3^2 = 0\},$$
2088: and let $Aut(L_{\zz})$ be the set of linear transformations that
2089: leave $L_\zz$ invariant.
2090: %
2091: %
2092: % Lemma 7.3
2093: %
2094: 
2095: \begin{lemma}~\label{lemma53}
2096:  $Aut(L_\zz) =O(3, 1, \zz)$.
2097: \end{lemma}
2098: \pf It is clear that $O(3,1;\zz) \subseteq  Aut(L_\zz)$. 
2099: To show the other direction,
2100: let $\bU \in Aut(L_\zz)$. For any integer point $\bv \in L_\zz$,
2101: $U\bv \in L_\zz$. Therefore $(\bU \bv)^T Q_\sL (\bU \bv)=0$, i.e.,
2102: $\bv^T (\bU^T Q_{\sL} \bU) \bv=0$.
2103: It is easy to check that the only symmetric matrices $\bQ$ satisfying
2104: $\bv^T \bQ \bv=0$ for all $\bv \in L_\zz$ are of the form 
2105: $\bQ=diag[-a, a, a, a]$.
2106: Hence $\bU^T \bQ_\sL \bU=c^2 \bQ_\sL$, where $c=det(\bU)$.
2107: 
2108: Let
2109: \[
2110: \bX=\left[\begin{array}{cccc}
2111: 1 & 1 & 1 & 1 \\
2112: 1 & -1 & 0 & 0 \\
2113: 0 & 0 & 1 & 0 \\
2114: 0  & 0 & 0 & 1 \end{array} \right].
2115: \]
2116: Note that every column vector of $\bX$ is an integer point in $L_\zz$. Let
2117: $\bY=\bU \bX$, and $\bZ=\bU^{-1} \bX$. Since $\bU \in Aut(L_\zz)$,  
2118: every column of $\bY$
2119: and $\bZ$ is also an integer point in $L_\zz$.
2120: Therefore $\det(\bY)=c \det(\bX)=-2c$,
2121: and $c \det(\bZ)=\det(\bX)=-2$. For each $(y_0, y_1, y_2, y_3)^T \in L_\zz$,
2122: we have $y_0^2=y_1^2+y_2^2+y_3^2$. It follows that
2123:  either $y_0$, $y_1$, $y_2$, $y_3$ are
2124: all even, or $y_0$ and exactly one of $y_1$, $y_2$, $y_3$ are even. 
2125: In both cases,
2126: $\det(\bY)$ and $\det(\bZ)$ are even. This forces $c=\pm 1$. Hence
2127: $\bU^T \bQ_\sL \bU =\bQ_{\sL}$, i.e., $\bU \in O(3,1)$.
2128: 
2129: Since $\bU$ maps points  $(1, \pm1, 0, 0)^T$, $(1, 0, \pm1,
2130: 0)^T$, $(1, 0, 0, \pm1)^T$ to integer points, if  $(a, b, c, d)$ is a row
2131: of $\bU$, then $a\pm b$, $a\pm c$, $a\pm d \in \zz$.
2132: Thus there exist integers $a'$, $b'$, $c'$, $d'$ of the same parity
2133: such that $a=\frac{a'}{2}$, $b=\frac{b'}{2}$, $c=\frac{c'}{2}$, 
2134: $d=\frac{d'}{2}$.
2135: However, by Equation \eqref{6.2for-2}  $\bU \bQ_\sL \bU^T =\bQ_\sL$. 
2136: This implies
2137:  $-a^2+b^2+c^2+d^2=\pm1$. Therefore ${b'}^2+{c'}^2+{d'}^2\equiv {a'}^2$
2138: (mod 4). Hence $a'$, $b'$, $c'$, $d'$ must all be even, 
2139: which means  $a$, $b$, $c$, $d$
2140: are integers, and $\bU \in O(3,1; \zz)$. ~~\bsq
2141: %
2142: %
2143: % Lemma 7.4
2144: %
2145: 
2146: \begin{lemma} \label{lemma54}
2147:  Let  $\cal G$ be a group of linear transformations that preserves
2148: $L_\zz$. If the action is  transitive, and there exists
2149: a point $v \in L_\zz$ such that the stabilizer
2150: $\sS_v=\{ \bU \in Aut(L_\zz) | \bU v=v\} \subseteq {\cal G}$,
2151: then ${\cal G}=Aut(L_\zz)$.
2152: \end{lemma}
2153: \pf  Clearly $\sG \subseteq Aut(L_\zz)$.
2154: For any $\bP \in Aut(L_\zz)$, assume $\bP(v)=v'$. Since
2155: $\sG$ acts transitively on $L_\zz$, there exists $\bG \in \sG$ such
2156: that $\bG(v')=v$. That is,  $\bG \bP(v)=v$. So $\bG \bP 
2157: \in \sS_v \subseteq \sG$, and
2158: then $\bP \in \bG^{-1} \sG=\sG$. This proves $Aut(L_\zz) \subseteq
2159: \sG$.~~~\bsq
2160: 
2161: %
2162: %
2163: % Lemma 7.5
2164: %
2165: 
2166: 
2167: \begin{lemma} \label{lemma55}
2168:   $\tilde \Gamma = G= \bJ_0 O(3,1;\zz)\bJ_0^{-1}$.
2169: \end{lemma} 
2170:                 
2171: \pf
2172:     By Lemma \ref{lemma53}, it is  sufficient to prove that
2173: \beq \label{6.2for-3}
2174: \bJ_0^{-1} \tilde \Gamma \bJ_0=\bJ_0 \tilde \Gamma \bJ_0 =Aut(L_\zz).
2175: \eeq
2176: It is straightforward to check that $\bJ_0 \bS_i \bJ_0$, $\bJ_0 \bS^T_i \bJ_0$,
2177: $\bJ_0 \bP_{\sigma} \bJ_0$ and $\bJ_0 \bD \bJ_0$ are integer matrices. 
2178: In particular,
2179: \beqs
2180: \bJ_0 \bS_1 \bJ_0=\left[ \begin{array}{rrrr}
2181:           2 & -1 & -1 & -1 \\
2182:           1 &  0 & -1 & -1 \\
2183:           1 & -1 &  0 & -1 \\
2184:           1 & -1 & -1 &  0 \end{array} \right],
2185: \eeqs
2186: $\bJ_0 \bP_{34} \bJ_0 =\bP_{34}$ and $\bJ_0 \bD \bJ_0 =diag[1,-1,-1,-1]$. 
2187: Therefore
2188: $\bJ_0 \tilde \Gamma \bJ_0 \subseteq O(3,1;\ZZ) = Aut(L_\zz)$.
2189:  
2190:    For any integer point $(y_0, y_1, y_2, y_3) \in L_\zz$,
2191: $-y_0^2+y_1^2+y_2^2+y_3^2=0$ implies \\
2192: $y_0+y_1+y_2+y_3\equiv 0~ (\bmod~ 2)$.
2193: Hence $\bJ_0 (y_0, y_1, y_2, y_3)^T$ is integral. It follows that
2194: \[
2195: \bJ_0(L_\zz)=
2196: \{(a_1, a_2, a_3, a_4) \in \zz^4 ~:~ (a_1, a_2, a_3, a_4) 
2197: \text{ are curvatures of a Descartes configuration }\}.
2198: \]
2199: Then Theorem \ref{reduction}  implies that 
2200: $\bJ_0 \tilde \Gamma \bJ_0$ acts transitively
2201: on the integral Lorentz light  cone $L_\zz$, where $-\bI$ exchanges the total
2202: orientation of a point in $L_\zz$. 
2203:                                                                               
2204: We use Lemma \ref{lemma54} to prove equation \eqref{6.2for-3}
2205: with $\sG=J_0 \tilde \Gamma J_0$.
2206: Let $\bv :=(1, 1, 0,0) \in L_\zz$ and consider its stabilizer
2207: $$\sS_{\bv} :=\{ \bU \in O(3, 1;\zz) ~:~ \bU \bv=\bv, \ \bU^T \bQ_\sL \bU 
2208: = \bQ_\sL\}.$$
2209: Assume $\bU=(u_{i,j})_{i,j=1}^4$.
2210: Solving the equations
2211:       \beq
2212: \bU(1,1,0,0)^T=(1,1,0,0), \qquad
2213: \bU^T \bQ_\sL \bU= \bQ_\sL,
2214: \eeq
2215: we obtain the following linear and quadratic relations
2216: between the entries of $\bU$:
2217: \beqs
2218: u_{12}=1-u_{11},~~~~~ u_{13}&= & u_{23},~~~~~ u_{14}=u_{24}, \\
2219: u_{22}=1-u_{21},~~~~ u_{32} &= & -u_{31},~~~~ u_{42}=-u_{41},
2220: \eeqs
2221: and
2222: \beqs
2223: u_{33}^2+u_{43}^2=u_{34}^2+u_{44}^2=1,& &\ u_{33}u_{34}+u_{43}u_{44}=0, \\
2224: u_{13}=u_{31}u_{33}+u_{41}u_{43},  & &\ u_{14}=u_{31}u_{34}-u_{41}u_{44}, \\
2225: u_{21}=\frac{1}{2}(u_{31}^2+u_{41}^2).~~~
2226: \eeqs
2227:   It follows that the matrix $\bU$ can be expressed as
2228: \beq\label{form}
2229: \bU=\left( \begin{array}{cccc}
2230: 1+t & -t & gm+hn & km+ln \\
2231: t & 1-t & gm+hn & km+ln \\
2232: m & -m & g & k \\
2233: n & -n & h & l \end{array}\right),
2234: \eeq
2235: where $t=(m^2+n^2)/2$, $g^2+h^2=k^2+l^2=1$, and $gk+hl=0$.
2236: Since $g, h, k, l \in \zz$, we must have
2237:  $(g, h), (k, l) \in \{(\pm 1, 0), (0, \pm 1)
2238: \}$.
2239:                                    
2240: We can classify the matrices of the form \eqref{form} into four types,
2241: up to a possible  multiplication by $\bP_{34}$)
2242: , as follows.
2243: \beqs
2244: \text{Type I}: \ \left( \begin{array}{cccc}
2245:  1+t & -t & m & n\\
2246:  t & 1-t & m & n\\
2247:  m & -m & 1 & 0\\
2248:  n & -n & 0 & 1 \end{array}\right),
2249: \qquad
2250: \text{Type I\@I}: \ \left( \begin{array}{cccc}
2251:  1+t & -t & m & -n\\
2252:  t & 1-t & m & -n\\
2253:  m & -m & 1 & 0\\
2254:  n & -n & 0 & -1 \end{array}\right),
2255: \eeqs
2256: \beqs
2257: \text{Type  I\@I\@I}: \ \left( \begin{array}{cccc}
2258:  1+t & -t & -m & n\\
2259:  t & 1-t & -m & n\\
2260:  m & -m & -1 & 0\\
2261:  n & -n & 0 & 1 \end{array}\right),
2262: \qquad
2263: \text{Type I\@V}: \ \left( \begin{array}{cccc}
2264:  1+t & -t & -m & -n\\
2265:  t & 1-t & -m & -n\\
2266:  m & -m & -1 & 0\\
2267:  n & -n & 0 & -1 \end{array}\right),
2268: \eeqs
2269: where $t=(m^2+n^2)/2$ and $m,n, t \in \zz$.
2270:                                                                                 
2271: We denote  a matrix of type $X$ with parameters $m,n$ by $\bU(m,n; {\rm X})$.
2272: The following equations can be easily checked.
2273: \beqs
2274:  \bU (m,n; {\rm I}) \bU (k,l; {\rm I})&=& \bU (m+k, n+l; {\rm I}),\\
2275:  \bU (m,n;\text{I\@I})  \bU (k,l; {\rm I}) & = &
2276:          \bU (m+k, n-l; \text{I\@I}),\\
2277:  \bU (m,n; \text{I\@I\@I})  \bU (k,l; \text{I}) & = &
2278:         \bU (m-k, n+l; \text{I\@I\@I}),\\
2279:  \bU (m,n; \text{I\@V})  \bU (k,l; \text{I}) & = &
2280:          \bU (m-k, n-l; \text{I\@V}).
2281: \eeqs
2282: Also we have
2283: \[
2284:  \bU (1,1; {\rm I})^{-1} =\bU (-1,-1;{\rm I}), \qquad
2285:  \bU (1, -1; {\rm I})^{-1}= \bU (-1, 1; {\rm I}).
2286: \]
2287: Therefore the stabilizer $\sS_v$ is generated by $\bP_{34}$,
2288: $\bU (0,0; {\rm I})$, $\bU (0,0; \text{I\@I})$
2289: $\bU (0,0; \text{I\@I\@I})$, $\bU (0,0; \text{I\@V})$
2290: together with $\bA= \bU (1,1;{\rm I}), \bB=\bU (1, -1;{\rm  I})$.
2291:                                                                                 
2292: Note that $\bU (0,0; \text{I})=\bU (0,0; \text{I\@I})^2$,
2293: $ \bU (0,0; \text{I\@I\@I})=\bP_{34} \bU (0,0; \text{I\@I}) \bP_{34}$,
2294: and $\bU (0,0; \text{I\@V})=\bU (0,0; \text{I\@I})
2295: \bU (0,0; \text{I\@I\@I})$.
2296: Thus $\sS_v$ is generated by the matrices $\bP_{34}$, $\bA$, $\bB$, and
2297: $\bC=\bU (0,0; \text{I\@I})$.
2298: 
2299: Now  one can check that
2300: \beqs
2301: \bP_{34} & = & \bJ_0 \bP_{34} \bJ_0 ~\in \bJ_0 \tilde \Gamma \bJ_0, \\
2302: \bC & = & {\rm diag}[1,1,1,-1] =\bJ_0(-\bI \bP_{23} \bP_{14} \bD)\bJ_0 
2303: ~\in \bJ_0 \tilde \Gamma \bJ_0,  \\
2304: \bA & = & (\bJ_0 \bS_1 \bJ_0 ){\rm diag}[1,1,-1, -1]= 
2305: \bJ_0 (\bS_1 \bP_{12} \bP_{34}) \bJ_0
2306: ~\in \bJ_0 \tilde \Gamma \bJ_0,\\
2307: \bB & = & \big(\bJ_0 \bP_{34}  \bJ_0\big)  \bC \big(\bJ_0 \bS_2 \bJ_0  \big) \bC ~\in
2308:  \bJ_0 \tilde \Gamma \bJ_0.
2309: \eeqs
2310: Hence $\sS_v \subseteq  \bJ_0 \tilde \Gamma \bJ_0.$ By Lemma \ref{lemma54}
2311: $ \bJ_0 \tilde \Gamma \bJ_0=Aut(L_\zz)=O(3,1;\zz)$, or equivalently,
2312: $\tilde \Gamma =\bJ_0 O(3,1;\zz) \bJ_0^{-1}$. This finishes the proof.
2313: ~~~\bsq \\
2314: %
2315: % proof of thm. 7.1
2316: %
2317: %
2318: 
2319: \vspace{.3cm}
2320: \noindent {\bf Proof of Theorem~\ref{th51}.}
2321: Lemma \ref{lemma52} and Lemma \ref{lemma55} show that
2322: $\Gamma \leq Aut(Q_D, \zz) \leq G$. Hence $\Gamma =Aut(Q_D, \zz)$ since
2323: $\Gamma$ is a subgroup of $G$ of index 2 and
2324: the duality matrix $\bD$ is not in $Aut(Q_D, \zz)$. In other words,
2325: $Aut(Q_D, \zz)$ is generated by the super-Apollonian group $\sA^S$ and the
2326: finite group of order 48 generated by  the $4 \times 4$ permutation matrices
2327: and $\pm I$.
2328: The super-Apollonian group $\sA^S$ is a normal subgroup
2329: of $Aut(Q_D, \zz)$,
2330: since $\bP_\sigma \bS_i \bP_{\sigma^{-1}} =\bS_{\sigma(i)}$, and
2331: $\bP_\sigma \bS_i^T \bP_{\sigma^{-1}} =\bS_{\sigma(i)}^T$. The index is 48
2332: since $\sA^S \cap (\text{Perm}_4 \times \{ \pm \bI\})=\bI$, (c.f. \S5
2333: of Part I \cite{GLMWY11}).
2334:                                              
2335:                                    
2336: By Lemma \ref{lemma55}, the group $G$ is generated by $Aut(Q_D, \zz)$
2337: and $\bD$. Note that $\bD^2=I$ and $\bD \bS_i \bD =\bS^T_i$. 
2338: It follows that the super-Apollonian group is a normal  subgroup 
2339: of $G$ with index 96. ~~\bsq \\
2340: 
2341: The second part of Theorem~\ref{th51} can be rephrased
2342: as asserting that the conjugate group
2343: $\tilde{\sA}^{S} = \bJ_0 \sA \bJ_0$ is a normal subgroup of
2344: index $96$ in $O(3,1)$. Its generators 
2345: $\tilde{\bS}_j = \bJ_0 \bS_j \bJ_0^{-1}$
2346: and $\tilde{\bS}_j^{\perp} = \bJ_0 \bS_j^{\perp} \bJ_0^{-1}$
2347: are given by
2348: $$
2349: \tilde{\bS}_1 
2350: \left[ \begin{array}{rrrr}
2351: 2 & -1 & -1 &-1 \\
2352: 1 & 0 & -1 & -1 \\
2353: 1 & -1 & 0 & -1 \\
2354: 1 & -1 & -1 & 0
2355: \end{array} 
2356: \right] ~~~\mbox{and}~~
2357: \tilde{\bS}_2  = \left[ \begin{array}{rrrr}
2358: 2 & -1 & 1 &1 \\
2359: 1 & 0 & 1 & 1 \\
2360: -1 & 1 & 0 & -1 \\
2361: -1 & 1 & -1 & 0
2362: \end{array}
2363: \right] 
2364: $$
2365: and
2366: $$
2367: \tilde{\bS}_3 = \left[ \begin{array}{rrrr}
2368: 2 & 1 & -1 &1 \\
2369: -1 & 0 & 1 & -1 \\
2370: 1 & 1 & 0 & 1 \\
2371: -1 & -1 & 1 & 0
2372: \end{array} 
2373: \right] ~~~\mbox{and}~~
2374: \tilde{\bS}_4  = \left[ \begin{array}{rrrr}
2375: 2 & 1 & 1 &-1 \\
2376: -1 & 0 & -1 & 1 \\
2377: -1 & -1 & 0 & 1 \\
2378: 1 & 1 & 1 & 0
2379: \end{array}
2380: \right] 
2381: $$
2382: The generators $\tilde{\bS}_j^{\perp} = \bJ_0 \bS_j^{\perp} \bJ_0^{-1}$
2383: are given by 
2384: $$
2385: \tilde{\bS}_j^{\perp} = (\tilde{\bS}_j)^{T},
2386: $$
2387: which follows using $\bS_j^{\perp} = \bS_j^T$ 
2388: and $\bJ_0 = \bJ_0^T = \bJ_0^{-1}.$
2389: 
2390: 
2391: 
2392: %*****************************************************************************
2393: %
2394: % SECTION 8. Super-integral Configurations
2395: %
2396: %*****************************************************************************
2397: %
2398: \section{Super-Integral Super-Packings}
2399: \setcounter{equation}{0}
2400: 
2401: This section  treats the strongest form of integrality for
2402: super-packings, which is that where one
2403: (and hence all) Descartes configurations in the
2404: super-packing have 
2405: an integral augmented curvature-center coordinate
2406: matrix $\bW_{\sD}$.
2407: We say that a Descartes configuration with this
2408: property is {\em super-integral}, and the
2409: same for the induced super-Apollonian packing.
2410: The following result classifies such packings.
2411: %
2412: %
2413: % Theorem 8.1
2414: %
2415: %
2416: 
2417: \begin{theorem}~\label{th81n}
2418: (1) These are  exactly $14$ different geometric
2419: super-packings that are super-integral.
2420: 
2421: (2) The set of ordered, oriented Descartes
2422: configurations that are super-integral comprise
2423: $672$ orbits of the super-Apollonian
2424: group.
2425: \end{theorem}
2426: 
2427: These packings are classified  here as rigid objects, not movable
2428: by Euclidean motions. 
2429: To prove this result, it suffices to determine which strongly-integral
2430: configurations are super-integral.
2431: The next result classifies 
2432: the possible types of super-integral Descartes configurations,
2433: according to the allowed value of their divisors.
2434: 
2435: %
2436: %
2437: % Theorem 8.2
2438: %
2439: %
2440: 
2441: \begin{theorem}~\label{super-int}
2442: Suppose that
2443: an ordered, oriented  Descartes configuration $\sD$ in $\RR^2$ has integral 
2444: curvature-center
2445: coordinates $\bM=\bM_{\sD}$, and let $g= \gcd(a_1, a_2, a_3, a_4)$,
2446: where $(a_1, a_2, a_3, a_4)$ is its first column of signed
2447: curvatures.
2448: Then $\sD$ has integral augmented curvature-center coordinates
2449: $\bW_{\sD}$ if and only if one of the following conditions hold. 
2450: 
2451: (i) $g=1,$ or
2452: 
2453: (ii) $g= 2$, and each row of $\bM$ has 
2454: the sum of its last two entries being $1~(\bmod~2)$. 
2455: 
2456: (iii) $g=4$, and the last columns rows of $\bM$ are congruent to
2457: $$
2458: \left[ \begin{array}{cc}
2459: 1 & 0 \\
2460: 1 & 0 \\
2461: 1 & 0 \\
2462: 1 & 0 \\ \end{array} \right]  ~~~\mbox{or}~~~
2463: \left[ \begin{array}{cc}
2464: 0 & 1 \\
2465: 0 & 1 \\
2466: 0 & 1 \\
2467: 0 & 1 \\ \end{array} \right] ~~(\bmod~2).
2468: $$
2469: \end{theorem}
2470: \pf
2471: By Theorem \ref{normal_form}, there exists a matrix $\bU \in \sA^S$ and 
2472: a permutation matrix $\bP$ such that 
2473: \[
2474: \bP \bU \bM_\sD =A_{m,n}[g] \text{ or } B_{m,n}[g],
2475: \]
2476: for some $m, n \in \{0,1\}$, 
2477: where $A_{m,n}[g]$ and $B_{m,n}[g]$ are given in Formula \eqref{ABmn}, 
2478: and their
2479: corresponding augmented curvature-center coordinate matrices are
2480: $\tilde A_{m,n}[g]$, $\tilde B_{m,n}[g]$, given in Formula \eqref{aug-ABmn}. 
2481: Since each generator of $\sA^S$ preserves the super-integrality, as
2482: well as the parity of every element of $\bW_\sD$,  it follows that 
2483: $\bW_\sD$ is integral if and only if one of the following conditions
2484: holds:
2485: \begin{enumerate}
2486: \item $g=1$, or
2487: \item $g=2$, and $m^2+n^2-1 \equiv 0~(\bmod~2)$, or
2488: \item $g=4$, and the reduced form is $A_{0,1}[4]$ or $B_{1,0}[4]$, up 
2489: to a permutation of rows.  
2490: \end{enumerate}
2491: In Case 2 we need $m+n \equiv 1 (\bmod 2)$; in Case 3 the condition is
2492: equivalent to the one stated in the theorem. 
2493: ~~~$\bsq$
2494: 
2495: 
2496: \paragraph{Proof of Theorem~\ref{th81n}.} 
2497: (1) Since each geometric super-packing corresponds to 48 distinct orbits of
2498: the super-Apollonian group on ordered, oriented Descartes configurations, 
2499: to show  there are exactly 14 geometric super-packings,
2500: it suffices to show there are exactly 672 orbits of the super-Apollonian
2501: group that 
2502: are super-integral, which is (2).
2503: 
2504: 
2505: 
2506: 
2507: (2)Theorems \ref{normal_form} and \ref{super-int} allow us to classify the 
2508: set of ordered, oriented Descartes configurations that are super-integral 
2509: by the action of $\sA^S$.  
2510: From the criterion of Theorem \ref{super-int}, we have:
2511: \begin{enumerate}
2512: \item For $g=1$, any strongly integral Descartes configuration is 
2513: super-integral. 
2514: \item For $g=2$, half of those orbits are super-integral, namely, those
2515:  whose reduced forms are $A_{0,1}[2], A_{1,0}[2], B_{0,1}[2]$ or $B_{1,0}[2]$, 
2516: up to 
2517:  a permutation of rows.
2518: \item For $g=4$, one fourth of those orbits are super-integral, namely, 
2519: those whose reduced form are $A_{0,1}[4]$ or $B_{1,0}[4]$, up to a permutation 
2520: of rows. 
2521: \item For $g \neq 1, 2, 4$, there are no super-integral Descartes 
2522: configurations.   
2523: \end{enumerate}
2524: 
2525: 
2526: 
2527: \begin{table}[htpb] 
2528: \begin{center}
2529: \begin{tabular}{|c|c|c|} 
2530: \hline 
2531: $\bg$  & \# of orbits of $\sA^S$  & Representative \\
2532: \hline
2533: $(1,1,1,1)$   & 96   & $A_{1,1}[1],~~~ B_{1,1}[1]$ \\
2534: $(2,1,1,1)$   & 96   & $A_{1,0}[1], ~~~B_{0,1}[1]$ \\
2535: $(1,1,2,1)$   & 48   & $A_{0,0}[1]$              \\ 
2536: $(1,1,1,2)$   & 48   & $B_{0,0}[1]$  \\
2537: $(4,1,2,1)$    & 48   & $A_{0,1}[1]$  \\
2538: $(4,1,1,2)$    & 48   & $B_{1,0}[1]$  \\ \hline 
2539: $(1,2,1,1)$    & 96   & $A_{1,0}[2], ~~~B_{0,1}[2]$ \\
2540: $(2,2,2,1)$    & 48    & $A_{0,1}[2]$  \\
2541: $(2,2,1,2)$    & 48     & $B_{1,0}[2]$  \\ \hline
2542: $(1,4,2,1)$    & 48     & $A_{0,1}[4]$  \\
2543: $(1,4,1,2)$    & 48     &  $B_{1,0}[4]$ \\
2544: \hline
2545: \end{tabular}
2546: \caption{Orbits of super-integral Descartes configurations classified by $\bg$.}
2547: ~\label{ta81}
2548: \end{center}
2549: \end{table} 
2550: 
2551: More details of this calculation are given in Table~\ref{ta81}.
2552: To explain the notation in Table~\ref{ta81}, 
2553: for any $4\times 4$ integral matrix $\bW$, let $g_i$ be the greatest common 
2554: divisor of entries $w_{1,i}$, $w_{2,i}$, $w_{3,i}$, $w_{4,i}$ 
2555: in the $i$-th column. 
2556: Then the action of $\sA^S$ preserves the vector $\bg=(g_1, g_2, g_3, g_4)$. 
2557: (For $\bW=\bW_\sD$, $g_2$ is the greatest common divisor of the curvatures.)  
2558: In each  row of the table  we give the number of orbits of $\sA^S$ formed by 
2559: the set of ordered, oriented  Descartes configurations that are super-integral
2560: with the given $\bg$. We also list the representatives of those orbits. 
2561: Note that each entry in the column labelled ``Representative'' 
2562: stands for 48 orbits, obtained 
2563: by taking two choices of (total) orientation, and 24 choices of permutation
2564: of rows. Also note the symmetry that the 
2565: configurations with $\bg$  are inverse of the ones with
2566: $\bg'=\bP_{12} \bg$.
2567: We  conclude from the table that the set of  ordered, oriented
2568: Descartes configurations that are super-integral comprise  
2569: $384(1+\frac{1}{2} +\frac{1}{4})=672$ orbits of the super-Apollonian group.
2570: ~~~$\bsq$
2571: 
2572: 
2573: 
2574: %****************************************************************************
2575: %
2576: % SECTION 9  Concluding remarks ?
2577: %
2578: %*****************************************************************************
2579: 
2580: \section{Concluding remarks}
2581: \setcounter{equation}{0}
2582: 
2583: This paper showed that the ensemble
2584: of all primitive,
2585: strongly integral Apollonian circle packings can  be
2586: simultaneously described 
2587: in terms of  an orbit of a larger discrete group,
2588: the super-Apollonian group, acting on the
2589: standard strongly-integral super-packing.  
2590: Study of the  locations of
2591: the individual integer packings inside 
2592: the standard  super-packing, presented in \S6,
2593:  leads  to interesting
2594: questions, not all of which are resolved.
2595: The standard super-packing also played a
2596: role in analyzing the structure of the
2597: super-Apollonian group as a discrete subgroup,
2598: carried out in  \S7.
2599: 
2600: The various illustrations show 
2601:  the usefulness of graphical representations, 
2602: as a guide to both finding and illustrating
2603: results. This contribution  is due in large part
2604: to the statistician co-authors (CLM and AW).
2605: Graphics were  particularly useful in finding extra
2606: symmetries of these objects, such as those illustrated
2607: in Figures 9--11 and subsequently proved by
2608: S. Northshield \cite{No03}.
2609: However one must not forget the
2610: adage of H. M. Stark \cite[p. 225]{St78}:
2611: ``A picture is worth a thousand words, provided 
2612: one uses another thousand words to justify
2613: the picture.''  Section 3 of this paper 
2614: provides such a justification for certain
2615: features of Figure 4.
2616: 
2617: There remain some open questions, particularly
2618: concerning the classification of all 
2619: integer root quadruples 
2620:  classified by fixed curvature $-N$ of the
2621: bounding circle. This quantity is known
2622: to be interpretable as a class number,
2623: as described in \cite[Theorem 4.2]{GLMWY2}.
2624: In \S6 we observed some symmetries of these
2625: root quadruples inside the standard super-packing,
2626: see Figures ~\ref{fig4}--\ref{fig6}.  
2627: There is a new invariant
2628: that can be associated to such quadruples,
2629: which is their nesting depth as defined in \S4
2630: with respect to the generating quadruple
2631: $\sD_0$ of the standard super-packing. It would
2632: be interesting to see whether this invariant
2633: might give some  further insight into class numbers. 
2634: 
2635:  
2636: 
2637: %****************************************************************************
2638: %
2639: % SECTION 10  Appendix: Strong Integrality Criterion 
2640: %
2641: %*****************************************************************************
2642: \newpage
2643: \section{Appendix.  Strong Integrality Criterion}
2644: % for  Descartes Configurations}
2645: \setcounter{equation}{0}
2646: 
2647: This appendix  establishes that to
2648: show a Descartes configuration is strongly integral
2649: it suffices to show that
2650:  three of its four circles are strongly integral.
2651: This affirmatively answers a question posed to us by K. Stephenson.
2652: 
2653: %
2654: %
2655: % Theorem 10.1
2656: %
2657: %
2658: 
2659: \begin{theorem}~\label{th101}
2660: A Descartes configuration $\sD$ has integral curvature-center
2661: coordinates $\bM_{\sD}$ if and only if it contains three
2662: circles having integer curvatures and whose
2663: curvature$\times$centers, viewed as complex numbers,  lie in $\zz[i]$.
2664: \end{theorem}
2665:                                                                                 
2666: \pf
2667: The condition is clearly necessary, and the problem is
2668: to show it is sufficient. We write the circle
2669: centers as complex numbers $\bz_j = x_j + i y_j$. So suppose $\sD$
2670: contains three circles with curvatures
2671: $b_1,~b_2, ~b_3 \in \zz$ and with
2672: curvature$\times$centers $b_1\bz_1,~b_2\bz_2,~b_3\bz_3 \in \zz[i]$.
2673: We must show that
2674: the fourth circle in the configuration has
2675: $b_4 \in \zz$ and $b_4\bz_4 \in \zz[i].$
2676: 
2677: For later use, we note that Theorem 3.1 of Part I \cite{GLMWY11} has a nice
2678: interpretation using complex numbers $\bz$
2679: to represent circle centers. This was formulated  in \cite{LMW02}
2680: as the Complex Descartes theorem. 
2681: It gives 
2682: \beql{901}
2683: b_1^2\bz_1^2 +  b_2^2\bz_2^2 + b_3^2\bz_3^2 +  b_4^2\bz_4^2 =
2684: \frac{1}{2}( b_1\bz_1 +  b_2\bz_2 + b_3\bz_3 +  b_4\bz_4)^2.
2685: \eeq
2686: and
2687: \beql{soddy}
2688: b_1^2\bz_1 +  b_2^2\bz_2 + b_3^2\bz_3 +  b_4^2\bz_4
2689: = \frac{1}{2} ( b_1\bz_1 +  b_2\bz_2 + b_3\bz_3 +  b_4\bz_4)
2690: (b_1 + b_2 +b_3 +b_4).
2691: \eeq
2692:                                                                                 
2693:   We claim  that $b_4 \in \zz.$ This is proved in
2694: the following two cases.
2695: 
2696: {\em Case 1. $b_1b_2b_3\neq 0$.}
2697:                                                                                 
2698: We first suppose
2699: that  $\bz_1=0$.
2700: If both $x_2$ and $x_3$ are zero, then $-\frac{1}{b_1} =\frac{1}{b_2}+\frac{1}{b_3}$, which means $b_1b_2+b_2b_3+b_3b_1=0$. Hence $b_4=b_1+b_2+b_3 \in \zz$.
2701: Otherwise by permuting $\bz_2$ and $\bz_3$ if
2702: necessary we may assume that $x_2 \ne 0$.
2703: Then the following equations encode the distance between 
2704: the circle centers, since the circles touch.
2705: \begin{eqnarray}
2706: x_2^2+y_2^2=(\frac{1}{b_1} + \frac{1}{b_2})^2, \label{xy1_sqr} \\
2707: x_3^2+y_3^2=(\frac{1}{b_1} + \frac{1}{b_3})^2, \label{xy_sqr} \\
2708: (x_3-x_2)^2+(y_3 - y_2)^2
2709: =(\frac{1}{b_2}+\frac{1}{b_3})^2. \label{xyxy1}
2710: \end{eqnarray}
2711: We wish to solve these equations for $y_3$ in terms of
2712: $b_1, b_2, b_3, x_1$ and $x_2$.
2713: To this end we subtract the
2714: first two equations from the third and obtain
2715: $$
2716: 2(x_2x_3 + y_2 y_3) = (\frac{1}{b_1} + \frac{1}{b_2})^2
2717: + (\frac{1}{b_1} + \frac{1}{b_3})^2 -(\frac{1}{b_2}+\frac{1}{b_3})^2 := R.
2718: $$
2719: Calling the right side of this equation $R$, we obtain
2720: $$
2721: x_3 = \frac{1}{2x_2}\left( R - 2y_2 y_3 \right).
2722: $$
2723: where the division  is allowed since $x_2 \ne 0$.
2724: Substituting this in the second equation yields a quadratic
2725: equation in $y_3$, with $x_3$ eliminated, namely
2726: (after multiplying by $4x_2^2$),
2727: $$
2728: 4(x_2^2 + y_2^2)  y_3^2 - 4R~ y_2y_3+R^2
2729: - 4x_2^2  (\frac{1}{b_1} + \frac{1}{b_3})^2 = 0.
2730: $$
2731: Since $y_3$ is
2732: rational, this  equation has rational  solutions, so the discriminant $\Delta$
2733: must be the square of a rational number. After some
2734: calculation one obtains
2735: \[
2736: \Delta=16^2 x_2^2 \left(\frac{ b_1b_2+b_1b_3+b_2b_3}{ (b_1b_2b_3)^2} \right).
2737: \]
2738: Since $x_2, b_1, b_2$ are
2739: nonzero it  follows that  $b_1b_2+b_2b_3+b_1b_3$ is a perfect square.
2740: Viewing  the Descartes equation as a quadratic equation in $b_4$
2741: we obtain the formula
2742: $b_4 :=b_1+b_2+b_3\pm2\sqrt{b_1b_2+b_2b_3+b_1b_3},$
2743: which shows that both roots are integers.
2744: These are the oriented curvatures of the
2745: two possible choices for the fourth circle in the
2746: Descartes configuration, so  $b_4 \in \ZZ$.
2747:                                                                                 
2748: Now assume that   $\bz_1= x_1 + i y_1$ is arbitrary. Define
2749: \begin{equation*}
2750: (s_2, t_2) := (x_2 - x_1, y_2 - y_1) ~~\mbox{and}~~ (s_3, t_3) := (x_3 - x_1,
2751: y_3-y_1).
2752: \end{equation*}
2753: Then $s_2, t_2, s_3, t_3$ are rational numbers, and they also satisfy
2754: equations \eqref{xy1_sqr}, \eqref{xy_sqr}, \eqref{xyxy1}. (just replace
2755: $x_2, y_2, x_3, y_3$ in \eqref{xy1_sqr}, \eqref{xy_sqr},
2756: \eqref{xyxy1} by $s_2,t_2,s_3, t_3$, respectively.)
2757: By the preceding argument, we again have
2758: $16s_2^2 (b_1b_2+b_2b_3+b_1b_3)/(b_1b_2b_3)^2 = q^2$
2759: for some rational number $q$.
2760: This implies that $b_1b_2+b_2b_3+b_1b_3$ is a perfect square and
2761: consequently that
2762: $b_4$ is an integer.
2763:                                                                                 
2764: {\em Case 2. $b_1b_2b_3=0$.}
2765: This  is proved similarly,  with an
2766: easier calculation.
2767:                                                                                 
2768: The claim follows, so $b_4 \in \ZZ$.
2769: 
2770: We now proceed to  show that
2771: $b_4\bz_4 \in \zz[i]$. Now  \eqn{901}  gives
2772: \begin{equation}~\label{reduce0}
2773: b_4\bz_4=b_1\bz_1+b_2\bz_2+
2774: b_3\bz_3\pm2\sqrt{b_1\bz_1b_2\bz_2+b_2\bz_2b_3\bz_3+b_1\bz_1b_3\bz_3}.
2775: \end{equation}
2776: The equation  \eqn{soddy} gives
2777: \begin{equation}~\label{reduce}
2778: (b_1+b_2+b_3-b_4)b_4\bz_4=2(b_1^2\bz_1+b_2^2\bz_2+b_3^2\bz_3)-(b_1+b_2+b_3+b_4)
2779: (b_1\bz_1+b_2\bz_2+b_3\bz_3).
2780: \end{equation}
2781: We treat two mutually exhaustive cases.
2782:                                                                                 
2783: {\em Case 1.  $b_1+b_2+b_3 \neq b_4.$}
2784:                                                                                 
2785: Then \eqn{reduce} gives
2786: $b_4\bz_4=x_4+iy_4$ for some rational
2787: numbers $x_4, y_4$. However
2788:   $b_1\bz_1, b_2\bz_2, b_3\bz_3$ are integers by hypothesis,
2789: whence \eqn{reduce0} shows that $b_4z_4$ is an algebraic integer.
2790: Since $x_4, y_4$  are rational, we conclude that
2791: $b_4\bz_4$ must be an integer, i.e., in $\zz[i]$.
2792:                                                                                 
2793: {\em Case 2.  $b_1+b_2+b_3=b_4$.}
2794:                                                                                 In this case, we have
2795: $b_1b_2+b_2b_3+b_1b_3 =0$. Now from
2796: \eqref{reduce}, we have
2797: \[
2798:  b_1^2\bz_1+b_2^2\bz_2+b_3^2\bz_3=(b_1+b_2+b_3)(b_1\bz_1+b_2\bz_2+b_3\bz_3),
2799: \]
2800: which can be simplified to
2801: \[
2802: b_2b_3\bz_1+b_1b_3\bz_2+b_1b_2\bz_3=0.
2803: \]
2804: Thus
2805: \begin{eqnarray*}
2806: (b_1\bz_1)(b_2\bz_2)+(b_2\bz_2)(b_3\bz_3)+(b_3\bz_3)(b_1\bz_1) & = &
2807: b_1(b_2\bz_2+b_3\bz_3)\frac{b_1b_3\bz_2+b_1b_2\bz_3}{-b_2b_3}
2808: +(b_2\bz_2)(b_3\bz_3) \\
2809: &=&-b_1^2\bz_2^2-b_1^2\bz_3^2+\bz_2\bz_3 \left(b_2b_3-
2810: \frac{b_1^2}{b_2b_3}(b_2^2+b_3^2)\right) \\
2811: &=& -b_1^2(\bz_2-\bz_3)^2.
2812: \end{eqnarray*}
2813: Now $b_1(\bz_2-\bz_3)$ is a Gaussian
2814: rational number whose square is integral.
2815: Hence $b_1(\bz_2-\bz_3)$ must be integral.
2816: It follows that $b_4\bz_4=b_1\bz_1+b_2\bz_2+b_3\bz_3\pm 2b_1(\bz_2-\bz_3)i$ is
2817: in $\zz[i]$.~~~$\bsq$ \\
2818:                                                                                 
2819: Since the Apollonian group consists of integer matrices,
2820: all Descartes configurations
2821: in an  Apollonian packing generated by  a strongly integral
2822: Descartes configuration are  strongly
2823: integral.  This explains
2824: the integrality properties of curvatures and
2825: curvature $\times$ center pictured in the packing
2826: in \S1, for example. The previous result now  gives
2827: a weaker necessary and sufficient condition for an Apollonian
2828: packing to be strongly integral.
2829: 
2830: %
2831: % Theorem 10.2
2832: %
2833: 
2834: \begin{theorem} \label{th32} 
2835: An Apollonian circle packing is strongly integral if
2836: and only if it contains three mutually tangent
2837: circles which have integer curvature-center
2838: coordinates.
2839: \end{theorem}
2840:                                                                                 
2841: \pf
2842: Suppose we are give three mutually tangent circles
2843: in the packing that are strongly integral.
2844: Any set of three mutually tangent circles in
2845: the packing is part of some Descartes configuration
2846: in this packing. This follows from the recursive
2847: construction of the packing, which has a finite
2848: number of circles at each iteration. If iteration $j$
2849: is the first iteration at which all three tangent
2850: circles are present, at that iteration they
2851: necessarily belong to a unique Descartes configuration.
2852: Theorem~\ref{th101} now implies that this
2853: Descartes configuration is strongly integral.
2854: It then follows that the whole Apollonian
2855: packing is strongly integral. This proves the
2856: ``if'' direction, and the ``only if'' direction
2857: is immediate.
2858: ~~~$\bsq$
2859:                                                                                 
2860: 
2861: 
2862: %*****************************************************************************
2863: %
2864: %  References 
2865: %
2866: %*****************************************************************************
2867: % 
2868: 
2869: 
2870: \clearpage
2871: \begin{thebibliography}{99}
2872: 
2873: \bibitem{AS97}
2874: D. Aharonov and K. Stephenson,
2875: Geometric sequences of discs in the
2876: Apollonian packing.
2877: Algebra i Analiz {\bf 9} (1997), No. 3, 104--140.
2878: [English version: St. Petersburg Math. J. {\bf 9} (1998), 509--545.]
2879: 
2880: \bibitem{Be83}
2881: A. F. Beardon,
2882: {\em The Geometry of Discrete Groups,}
2883: Springer-Verlag: New York 1983.
2884: 
2885: \bibitem{GLMWY11}
2886: R. L. Graham, J. C. Lagarias, C. L. Mallows, A. Wilks and C. Yan,
2887: Apollonian circle packings: geometry and group theory
2888: I. The Apollonian group, preprint.
2889: eprint: {\tt arXiv math.MG/0010298}.
2890: 
2891: \bibitem{GLMWY13}
2892: R. L. Graham, J. C. Lagarias, C. L. Mallows, A. Wilks and C. Yan,
2893: Apollonian circle packings: geometry and group theory
2894: III. Higher dimensions, 
2895: eprint: {\tt arXiv math.MG/0010324}.
2896: 
2897: \bibitem{GLMWY2}
2898: R. L. Graham, J. C. Lagarias, C. L. Mallows, A. Wilks and C. Yan,
2899: Apollonian circle packings: number theory, 
2900: Journal of Number Theory, {\bf 100} no.1 (2003), 1--45. 
2901: [eprint: arXiv math.NT/0009113]
2902: 
2903: \bibitem{LMW02}
2904: J. C. Lagarias, C. L. Mallows and A. Wilks,
2905: Beyond the  Descartes circle theorem,
2906: Amer. Math. Monthly {\bf 109} (2002),  338--361.
2907: [eprint: {\tt arXiv math.MG/0101066}]
2908: 
2909: \bibitem{M82}
2910: B. B. Mandelbrot,
2911: {\em The Fractal Geometry of Nature},
2912: Freeman: New York, 1982.
2913: 
2914: 
2915: 
2916: \bibitem{Max82}
2917: G. Maxwell,
2918: Sphere packings and hyperbolic reflection groups.
2919: J. Algebra {\bf 79} (1982), 78--97.
2920: 
2921: 
2922: \bibitem{MSW02}
2923: D. Mumford, C. Series, and D. Wright,
2924: {\em Indra's Pearls: The Vision of Felix Klein},
2925: Cambridge U. Press: Cambridge 2002.
2926: 
2927: \bibitem{No03}
2928: S. Northshield,
2929: On Apollonian circle packings,
2930: preprint.
2931: 
2932: 
2933: %\bibitem{Ru70}
2934: %W. R\"uhl,
2935: %{\em The Lorentz group and harmonic analysis},
2936: %W. A. Benjamin: New York 1970.
2937: 
2938: \bibitem{So92}
2939: B. S\"oderberg,
2940: Apollonian tiling, the Lorentz group, and regular trees.
2941: Phys. Rev. A {\bf 46} (1992), No. 4, 1859--1866.
2942: 
2943: \bibitem{St78}
2944: H. M. Stark,
2945: {\em An Introduction to Number Theory}, 
2946: M. I. T. Press: Cambridge-London 1978. 
2947: 
2948: \bibitem{Wi81}
2949: J. B. Wilker,
2950: Inversive Geometry,
2951: in: {\em The Geometric Vein}, (C. Davis, B. Gr\"unbaum, F. A. Sherk, Eds.),
2952: Springer-Verlag: New York 1981, pp. 379--442.
2953: 
2954: \end{thebibliography}
2955: 
2956: {\tt
2957: \begin{tabular}{lllll}
2958: email: & graham@ucsd.edu \\
2959: & lagarias@umich.edu \\
2960:   & colinm@research.avayalabs.com \\
2961:  & allan@research.att.com \\
2962: & cyan@math.tamu.edu
2963: \end{tabular}
2964:  }
2965: 
2966: 
2967: 
2968: \end{document}
2969: 
2970: 
2971: 
2972: 
2973: 
2974: 
2975: 
2976: 
2977: 
2978: 
2979: 
2980: 
2981: 
2982: 
2983: 
2984: 
2985: