math0310273/nor.tex
1: \documentclass[12pt]{amsart}
2: \usepackage{graphicx}
3: %\usepackage{tabularx}
4: 
5: \newcommand{\kt}{K_t}
6: \newcommand{\kr}{K_r}
7: \newcommand{\ym}{\mathcal{YM}}
8:  
9: \newcommand{\lcr}{\raisebox{-5pt}{\mbox{}\hspace{1pt}
10:                  \includegraphics{leftcross.eps}\hspace{1pt}\mbox{}}}
11: \newcommand{\ift}{\raisebox{-5pt}{\mbox{}\hspace{1pt}
12:                  \includegraphics{infinity.eps}\hspace{1pt}\mbox{}}}
13: \newcommand{\zer}{\raisebox{-5pt}{\mbox{}\hspace{1pt}
14:                  \includegraphics{zero.eps}\hspace{1pt}\mbox{}}}
15: \newcommand{\rkin}{\raisebox{-7pt}{\mbox{}\hspace{1pt}
16:                  \includegraphics{rightkink.eps}\hspace{1pt}\mbox{}}}
17: \newcommand{\lkin}{\raisebox{-7pt}{\mbox{}\hspace{1pt}
18:                  \includegraphics{leftkink.eps}\hspace{1pt}\mbox{}}}
19: \newcommand{\smst}[1]{\makebox[0pt]{\scriptsize{$#1$}}}
20: \newcommand{\displaystyl}{\displaystyle}
21: \newcommand{\Tet}{\mathrm{Tet}\begin{pmatrix}}
22: \newcommand{\shad}{\mathrm{Shad}}
23: \newcommand{\cS}{\mathcal{S}}
24: %\newcommand{\displaystyl}{}
25: \newtheorem{theorem}{Theorem}
26: \newtheorem{lemma}{Lemma}
27: \newtheorem{cor}{Corollary}
28: \newtheorem{prop}{Proposition}
29: \newtheorem{definition}{Definition}
30: \newtheorem{example}{Example}
31: \newtheorem{scholium}{Scholium}
32: \newtheorem{remark}{Remark}  
33: \newtheorem{fact}{Fact}
34: 
35: \setlength{\oddsidemargin}{18pt}
36: \setlength{\evensidemargin}{18pt}
37: \setlength{\textwidth}{435pt}
38: 
39: 
40: \title{The Quantum Content of the Normal Surfaces in a Three-Manifold}
41: 
42: \author{Charles Frohman}
43: 
44: \address{Department of Mathematics, University of Iowa, Iowa City, IA
45: 52242, USA}
46: 
47: \email{\tt frohman@math.uiowa.edu}
48: 
49: \author{Joanna Kania-Bartoszynska}
50: 
51: \address{Department of Mathematics, Boise State University, Boise, ID
52: 83725, USA} 
53: 
54: \email{\tt kania@math.boisestate.edu}
55: 
56: 
57: \thanks{This material is based upon work supported by the National
58: Science Foundation under Grant No.0207030  and under Grant No.0204627}
59: \keywords{Turaev-Viro invariants of $3$-manifolds, normal surface,
60: regular spine, quantum 6j-symbol.}
61: \subjclass{57M27}
62: \begin{document}
63: \begin{abstract}
64: 
65: The formula for the Turaev-Viro invariant of a $3$-manifold depends 
66: on a complex parameter $t$. When $t$ is not a root of unity, the formula 
67: becomes an infinite sum. This paper analyzes convergence of this sum  
68: when $t$ does not lie on the unit circle, in the presence of an efficient 
69: triangulation of the three-manifold. The terms of the sum can be indexed by
70: surfaces lying in the three-manifold. The contribution of a surface is 
71: largest when the surface is normal and when its genus is the lowest. 
72: 
73: \end{abstract}
74: \maketitle
75: 
76: \section{Introduction} This paper initiates the study of non-perturbative
77: quantum invariants of three-manifolds $M$  away from roots of unity. 
78: Turaev and Viro \cite{TV} defined  invariants of closed 3-manifolds
79: as  state sums depending on a complex parameter $t$. When $t$ is 
80: a root of unity this sum is finite.
81: At values of $t$ other than roots of unity the formula for
82: the Turaev-Viro invariant becomes an infinite sum. The partial sums
83: could oscillate wildly, so that even after
84: renormalizing the series does not converge. However, we are able to show 
85: that for a
86: special class of spines of some three-manifolds, the oscillation does not
87: occur and there is a limit. 
88: 
89: The key is
90: to see the invariant  as a sum over surfaces in the manifold.
91: An 
92: efficient ideal triangulation \cite{JR} of the manifold is one where
93: the only normal spheres and
94: tori are links of the boundary components of the manifold. Given an efficient
95: ideal triangulation we know what the invariant should be.
96: It is a sum over surfaces carried by a spine dual to the efficient
97: triangulation. 
98: The surfaces contribute to the sum in a way that fits the modern 
99: approach to
100: normal surface theory.
101: The study of normal surfaces \cite{JR} 
102: has been augmented by looking at surfaces that
103: aren't normal, and coming to an understanding of how a surface fails to be
104: normal. 
105: The farther a surface is from being normal, the less it contributes to
106: the sum for the invariant. The higher the genus of a surface the less it
107: contributes. In a very real sense, the Turaev-Viro invariant is a measure of
108: the normal surface theory of $M$.
109: 
110: 
111: In  section \ref{prel}  we broach preliminary concepts 
112: relating to special functions, and to spinal and normal surfaces in an
113: ideal triangulation of a $3$-manifold. This is followed by  section \ref{6jdet}
114: that studies  properties and limiting behavior 
115: of the  
116: $6j$-symbols. Section \ref{norandspi} is concerned with estimates  of
117: the contributions of spinal surfaces to the state sum.
118: The final section proves the result about the convergence of the
119: infinite state sum.
120: 
121: The authors would like to thank Ian Agol, William Jaco, Marc Lackenby, Sergei
122: Matveev, 
123: Dennis Roseman and Hyam Rubinstein for enlightening conversations about normal
124: surface theory.
125: 
126: 
127: \section{Preliminaries}\label{prel}
128: \subsection{Special Functions} The formulas in this section are taken
129: from \cite{KL}, however we use the variable $t$ instead of $A$.
130: Throughout this paper $t$ is a real number with $0<t<1$.
131: There are several functions of $t$  that we will work with. 
132: The first
133: is known as {\em quantized $n$},
134: \begin{equation} [n]=\frac{t^{2n}-t^{-2n}}{t^2-t^{-2}}.\end{equation}
135: The next is just a variation on the first,
136: \begin{equation}\Delta_n=(-1)^n\frac{t^{2n+2}-t^{-2n-2}}{t^2-t^{-2}}=(-1)^n[n+1].\end{equation}
137: There is quantized factorial, defined recursively by $[0]!=1$ and
138: \begin{equation} [n]!=[n][n-1]!.\end{equation}
139: 
140: A triple of nonnegative integers $(a,b,c)$ is admissible if their sum
141: is even and they satisfy every possible triangle inequality.
142: Admissibility is the necessary and sufficient condition for the
143: existence of  a Kauffman triad on $a$, $b$ and $c$.
144: Suppose that $(a,b,c)$ is admissible.
145: Arrange $a$ points, $b$ points and
146: $c$ points on the sides of a triangle.
147:  There exists a system of disjoint
148: proper arcs  joining opposite sides of
149: the triangle having those points as their boundary.
150: Figure \ref{adm} shows the admissible triple
151: $(2,3,3)$. 
152: \begin{figure}
153: \includegraphics{adm.eps}
154: \caption{Admissible triple $(2,3,3)$.}\label{adm}
155: \end{figure}
156: The number of strands running between the family of $a$ points, and the
157: family of $b$ points is $x_1=\frac{a+b-c}{2}$, between the $b$ points and the
158: $c$ points is $x_2=\frac{c+b-a}{2}$, and  between the $a$ points and the $c$
159: points is
160: $x_3=\frac{a+c-b}{2}$. Admissibility is equivalent to the statement that all
161: three functions are nonnegative and integral. We call $x_1$, $x_2$ and
162: $x_3$ the {\em strand numbers} of the triple $(a,b,c)$. 
163: 
164: You can always add two
165: admissible triples and the result will be admissible, but you cannot always
166: subtract them. However, 
167: 
168: \begin{prop}\label{subtract}
169: Suppose that $(a,b,c)$ and $(a',b',c')$ are admissible and
170: $x_i$ and $x'_i$ are the strand numbers corresponding to the two triples.
171: If $x_i\geq x'_i$ for all $i$, then  $(a-a',b-b',c-c')$ is
172: admissible. 
173: \end{prop}
174: 
175: \proof The strand numbers are linear functions of the triples, thus
176: the strand numbers for $(a-a',b-b',c-c')$ are integral and nonnegative.
177: \qed
178: 
179: 
180: If $(a,b,c)$ is an admissible triple, define, \cite{KL}
181: \begin{equation}
182: \theta(a,b,c)=(-1)^{\frac{a+b+c}{2}}\frac{[\frac{a+b+c}{2}+1]!
183: [\frac{a+b-c}{2}]![\frac{a+c-b}{2}]![\frac{b+c-a}{2}]!}{[a]![b]![c]!}.\end{equation}
184: 
185: 
186: 
187: 
188: Suppose that a tetrahedral net has been labeled as in  Figure \ref{tet} 
189: where the letters are nonnegative integers and the triples appearing at each
190: vertex $v$ are admissible. 
191: \begin{figure}
192: %\begin{picture}(56,120)
193: \raisebox{-8pt}{\includegraphics{tet1.eps}
194: \put(-50,16){$a$}
195: \put(-60,49){$b$}
196: \put(5,49){$e$}
197: \put(-23,70){$c$}
198: \put(-15,40){$d$}
199: \put(-40,49){$f$}}
200: %\end{picture}
201: \caption{Tetrahedral net}\label{tet}
202: \end{figure}
203: The tetrahedral coefficient \cite{KL,MV} is the quantity
204: \begin{equation}\text{Tet}\begin{pmatrix} a & b & e \\ c & d &
205: f\end{pmatrix}=\end{equation}
206: \[\frac{\prod_v[x_{v,1}]![x_{v,2}]![x_{v,3}]!}{[a]![b]![c]![d]![e]![f]!}
207: \sum_{s=m}^M
208: \frac{(-1)^s[s+1]!}{[B_1-s]![B_2-s]![B_3-s]![s-A_1]![s-A_2]![s-A_3]![s-A_4]!},
209: \]
210:  where the $B_i$ are half the sums of the labels over the four cycles, the
211:  $A_i$ are half the sums of the labels at each vertex, $m$ is the maximum of
212:  the $A_i$ and $M$ is the minimum of the $B_i$.
213: The $x_{v,i}$ are the
214:  strand numbers of the admissible triple at the vertex $v$.
215: 
216: The {\em unitary $6j$ symbol} \cite{Ro} is the quantity:
217: \begin{equation} \left\{\begin{matrix} a & b & e \\ c & d &
218: f\end{matrix}\right\}_u=\frac{\text{Tet}\begin{pmatrix} a & b & e \\ c & d &
219: f\end{pmatrix}}
220: {\sqrt{\theta(a,d,e)\theta(b,c,e)\theta(a,b,f)\theta(c,d,f)}}.\end{equation}
221: A consequence of admissibility is that the denominator is a square root of a
222: positive number, so the formula is unambiguous.
223: 
224: 
225: Letting $q=t^4$, let $(x;q)_n=\prod_{i=1}^n(1-xq^{i-1})$.
226: We need the following fact \cite{GR}:
227: The function
228: \begin{equation}(x;q)_{\infty}=\prod_{i=1}^{\infty} (1-xq^{i-1})\end{equation} 
229: is well defined when $|q|<1$. In particular, $(q;q)_{\infty}$
230: is well defined.  
231: Notice that 
232: \begin{equation}\label{Delta} 
233: [n]=t^{-2n+2}\frac{1-q^{n-1}}{1-q},
234: \end{equation}
235: and
236: \begin{equation}
237: \Delta_n=(-1)^nt^{-2n}\frac{1-q^n}{1-q},  
238: \end{equation}
239: so that
240: \begin{equation}\label{fact}
241: [n]! = t^{-(n-1)n} \displaystyle\frac{(q;q)_n}{(1-q)^n},
242: \end{equation}
243: and
244: \begin{equation}\theta(a,b,c)= (-1)^{\frac{a+b+c}{2}}\frac{t^{-a-b-c}}{1-q}\  
245: \frac{(q;q)_{\frac{a+b+c}{2}+1}(q;q)_{\frac{a+b-c}{2}}
246: (q;q)_{\frac{b+c-a}{2}}(q;q)_{\frac{a+c-b}{2}}}
247: {(q;q)_a(q;q)_b(q;q)_c}.
248: \end{equation}
249: 
250: The quantities $\Delta_n$, $\theta(a,b,c)$ and 
251: $\text{Tet}\begin{pmatrix} a & b & e \\ c & d &
252: f\end{pmatrix}$ can be understood as the Kauffman brackets of colored graphs
253: \cite{KL}.
254: 
255: \subsection{Spines and Ideal Triangulations}
256: An {\em ideal triangulation}  \cite{BP} of the compact three-manifold $M$ is a 
257: union of  tetrahedra
258: joined along faces with their vertices removed so that the result is
259: homeomorphic to the interior of $M$.  
260: 
261: A surface is {\em normal} with respect
262: to the triangulation \cite{JR} if it intersects each tetrahedron in 
263: triangles and quadrilaterals (quads) as in Figure \ref{norm1}.
264: \begin{figure}
265: \includegraphics{norm1.eps}
266: \caption{Normal surface intersecting a tetrahedron}\label{norm1}
267: \end{figure}
268: Parameterize normal surfaces by their intersection
269: with the edges of the tetrahedra. Arrange these numbers in a $2\times
270: 3$ array, so that  each column of the array is the number of points of 
271: intersection
272: of the normal surface with two opposite edges of the tetrahedron. More
273: specifically, there is
274: a tetrahedral net dual to the $1$-skeleton of the tetrahedron, lying on the
275: boundary of the tetrahedron, as pictured in Figure \ref{net}. 
276: The intersection of the normal surface with the boundary of the tetrahedron 
277: is a
278: family of circles carried by this net, the number of strands carried by an
279: edge of the net is the intersection number of the normal surface with the edge
280: of the tetrahedron transverse to the particular edge of the net. We form the
281: array of nonnegative integers just as if we were indexing a tetrahedral
282: coefficient, see Figure \ref{tet}, where the label on the edge is the number
283: of strands carried by that edge.  Let
284: $C_1,C_2,C_3$ be the sums of the columns of the array, named so
285: that
286: $C_1\geq C_2 \geq C_3$.
287: \begin{figure}
288: \includegraphics{net.eps}
289: \caption{Dual tetrahedral net}\label{net}
290: \end{figure}
291: 
292: \begin{prop}\label{normalsurface}
293:  An array
294: of nonnegative integers corresponds to a normal surface if and only if
295: the integers assigned to the three edges around each face of the
296: tetrahedron form an admissible triple, and 
297: $C_1=C_2$. 
298: \end{prop}
299: \proof Think of a face of a
300: tetrahedron as a 
301: triangle, and the intersection of the normal surface with the triangle as a
302: system of arcs joining points on the three edges.
303: As we are joining the points on the three sides of the triangle by
304: nonintersecting arcs, the triple around each face must be admissible.
305:  The second condition follows from the fact that the intersection of
306:  a normal surface with any tetrahedron can
307:  only contain triangles and one type of quad. \qed
308: 
309: You cannot necessarily add normal surfaces, because if
310: two normal surfaces have different quads in the same tetrahedron, their double
311: curve sum may  no longer be normal. However, there is always a finite family of
312: normal surfaces so that every normal surface can be written as an integral sum
313: of those surfaces, with nonnegative coefficients.
314: 
315: Letting $\mathbb{N}$ denote the nonnegative integers, a {\em rational
316: cone} is the solution of
317: a family of linear homogeneous equations with integer coefficients
318: in $\mathbb{N}^k$ for some $k$.
319: An element of a rational cone is {\em irreducible} if it cannot be
320: written  as the sum of two elements of the
321: cone in a nontrivial way. It is a classical result that a rational
322: cone has only finitely many irreducible elements and they generate the cone additively. The set of
323: irreducibles is called a {\em Hilbert basis} for the cone.
324: 
325: The normal surfaces form a rational cone. The class of
326: normal surfaces is a subset of a more general class of surfaces, the
327: {\em spinal surfaces}.
328: 
329: 
330: Let \begin{equation}X=\{(x,y,z)\in \mathbb{R}^3| z=0 \ \text{or} \ (z\geq 0 \ \text{and} \
331: x=0) \ \text{or} \ (z\leq 0 \ \text{and} \ y=0) \} .\end{equation}
332: A subset $Y$ of the 3-manifold $M$ is {\em modeled} on $X$ if 
333:  for every 
334: $p \in Y$ there is an open neighborhood  $U$ of $p$, open set
335: $V\subset\mathbb{R}^3$   and
336:  a homeomorphism $\phi: U \rightarrow V$, with
337: $\phi^{-1}(X)=Y \cap U$. There is a decomposition  of $Y$ into 
338: vertices, (open) edges and (open) faces  coming from the natural 
339: decomposition of 
340: $X$ into vertices, edges and faces. 
341: 
342: We say $S \subset M$ is a {\em regular spine} \cite{Ma, Pi} if:
343: \begin{enumerate}
344: \item $S$ is  modeled on $X$.
345: \item $M -S $ is homeomorphic to $\partial M \times [0,1)$.
346: \item $S$ has at least one vertex.
347: \item Every edge of $S$ has a vertex in its closure.
348: \item Every face of $S$ is simply connected.
349: \end{enumerate}
350: 
351: \begin{prop}
352:   Ideal triangulations and regular spines are in one to one correspondence up
353:   to isotopy via duality.
354: \end{prop}
355: \proof
356: For each ideal triangulation there is a regular spine. Put a vertex 
357: in the
358: center of each tetrahedron. Join the vertices in adjacent tetrahedra by edges,
359: and then form faces of the spine that intersect the edges of the triangulation
360: transversely and 
361: are bounded by the edges of the spine. 
362: The intersection of the spine with a tetrahedron is pictured in
363: Figure \ref{spine}.
364: \begin{figure}
365: \includegraphics{spine.eps}
366: \caption{Intersection of the spine with a tetrahedron}\label{spine}
367: \end{figure}
368: Similarly for each regular spine there is an ideal triangulation, so that its
369: six edges intersect the six faces of the spine coming into the vertex 
370: transversely,
371: and each edge intersects exactly one face.\qed
372: 
373: Given a spine $S$ of $M$ and a  simple closed curve 
374: $\kappa \subset \partial M$ there is a possibly singular  annulus
375: $A_{\kappa} \subset M$ having $\kappa$ as one boundary component so that the
376: intersection of $A_{\kappa}$ 
377: with $S$ is the other boundary component of $A_{\kappa}$. 
378: The annulus is constructed by taking the closure of the points lying
379: over $\kappa$ in the 
380: product structure on $M-S$. The singularities of $A_{\kappa}$ come from the
381: fact that the map from $\partial M$ to $S$ given by following the lines of the
382: product structure is two to one along faces.
383: Since there is some ambiguity in the product structure we
384: can choose the annulus $A_{\kappa}$ so that it is in general position with
385: respect to the spine. This means its boundary  misses the vertices of the
386: spine, intersects the edges transversely and its only singular points are
387: transverse double points  occurring in the interior of the faces of the
388: spine.  If $\mathcal{C}$ is a system of disjoint  simple closed curves in 
389: $\partial M$
390: let $A_{\kappa}$, where $\kappa \in C$, be a system of disjoint annuli
391: corresponding to the curves in $C$ that is in general position with respect to
392: the spine $S$. 
393: The union $S(C)=S \cup \left( \cup_{\kappa \in C} A_{\kappa}\right)$ 
394: is called 
395: {\em the augmentation of the spine with respect to $C$}. Except for points on
396: $C$ the augmentation is still modeled on $X$, so it can be decomposed
397: into vertices, edges and faces just as a spine. If the spine is regular
398: then the faces of the augmentation are simply connected except for the
399: annular faces  with one boundary component a curve in $C$.
400: 
401: An {\em admissible coloring} of a spine is an assignment of a nonnegative 
402: integer to each face of the spine so that the integers assigned to the 
403: three faces meeting along each edge form an admissible triple. Given an
404: admissible coloring of the spine there is a {\em spinal surface} built as
405: follows. If the face $f$ carries the integer $u_f$ then take $u_f$ parallel
406: copies of $f$. Along the edges glue the faces together so that they look like
407: the Cartesian product of a triple of arcs at a vertex with an interval. The
408: triple $(2,3,3)$ occurring along an edge is shown in Figure \ref{glue}.
409: \begin{figure}
410: \includegraphics{glue.eps}
411: \caption{Building a spinal surface}\label{glue}
412: \end{figure}
413: So far, the surface constructed intersects the boundary of a small ball at 
414: each vertex in a
415: collection of circles arranged along a tetrahedral net. To finish the
416: construction,  fill in the surface
417: inside each ball with a disk for each circle in the net.
418: 
419: Topologically, the spinal surfaces are those surfaces that intersect the
420: tetrahedra in disks, so that their intersection with any face of a tetrahedron
421: consists of arcs whose endpoints lie in distinct edges of the face.
422: The spinal surfaces form an additive cone, as the sum of two admissible
423: colorings is an admissible coloring. However, Euler characteristic is not
424: always additive under sum. Clearly, spinal surfaces are a larger class
425: than the normal surfaces associated to the dual triangulation.
426: We can identify the normal surfaces inside
427: the spinal surfaces by looking at the tetrahedral net at each vertex. In
428: specific at each vertex we can define the three column sums of the tetrahedral
429: net and order them so that $C_1 \geq C_2 \geq C_3$.
430: \begin{remark}
431: By  Proposition \ref{normalsurface} the surface is normal in the dual
432: ideal triangulation if and only if at each vertex
433: $C_1=C_2$. 
434: \end{remark}
435: 
436: 
437: The spinal surfaces form a rational cone. The proper domain
438: is the Cartesian product of copies of $\mathbb{N}$, one for each strand
439: number. The color on a face is the sum of the two adjacent strand
440: numbers along an edge of the face. The equations defining the cone
441: come from the 
442: requirement that the computed color of a face must be the same no
443: matter what edge of the face you compute it along. Thus we have the
444: following: 
445: \begin{fact}
446: There is a set of primitive spinal surfaces $\{F_i\}$ so that
447: every spinal surface can be written as a nonnegative  sum of the
448: $\{F_i\}$'s.
449: \end{fact}
450: Suppose now that $C$ is a system of disjoint simple closed curves in $\partial
451: M$ and $S(C)$ is an augmentation of the spine with respect to $C$. An
452: admissible coloring of $S(C)$ is defined the same way as an admissible coloring
453: of a spine except that the annular faces can only carry the color $1$. There
454: is once again a correspondence between admissible colorings and surfaces, but
455: now the surfaces have boundary equal to the union of the curves in $C$. 
456: 
457: In order to understand the Euler characteristic of a surface carried by a
458: spine or an augmented spine we need to understand how many circles there are
459: in a colored tetrahedral net.
460: 
461: \begin{prop}
462:   Suppose that a tetrahedral net has column sums $C_1 \geq C_2 \geq C_3$. 
463: The number of circles in the net is $\gcd{(C_1-C_2,C_1-C_3)}/2
464: +C_2+C_3-C_1 $. 
465: \end{prop}
466: 
467: %We need to estimate the Euler characteristic of a spinal surface.
468: 
469: %\begin{theorem}
470: %  If $F$ is a spinal surface, let $u_f$ be the color assigned to the face $f$
471:  % to define $F$, let $C_{v,1}\geq C_{v,2}\geq C_{v,3}$
472: %to be its column sums at each vertex, and let  If $x_F$ is the Euler characteristic of
473: %$F$ then 
474: %\[ x_F\geq \sum_f u_f -\sum_v \frac{1}{4}(C_{1,v}-C_{2,v})(C_{1,v}-C_{3,v})
475: %+C_{1,v}/2.\]
476: %\end{theorem}
477: 
478: %The theorem will follow from the following:
479: 
480: %\begin{lemma}
481:  % Given a tetrahedral net with curves running parallel along each edge so that
482:  % the column sums of the numbers of edges are $C_{1}\geq C_{2}\geq C_{3}$,
483: %then the number of circles in in the net is greater than or equal to
484: %\[ \frac{-1}{4}(C_1-C_2)(C_1-C_3)+\frac{1}{2}(C_2+C_3).\]
485: %\end{lemma}
486: 
487: \proof Unless
488: a tetrahedral net is of the form 
489: \begin{equation} 
490: \begin{pmatrix}
491:   a & b & a+b \\ a & b & a+b 
492: \end{pmatrix}\end{equation}
493: with $a$ and $b$ nonzero
494: then there is always a simple closed curve in the net that is the boundary of
495: one of the faces of the tetrahedron. 
496: 
497: %If you remove this from the net, the quadratic part of the
498: %right hand side is fixed and the linear part goes down by $1$. Hence the
499: %inequality is proved if we can establish it for nets like the one above.
500: %In this case the number of circles can be seen to be $\gcd{(a,b)}$. Hence
501: %we are trying to prove that 
502: %\[ \gcd{(a,b)}\geq \frac{-1}{4}(2a)(2b)+\frac{1}{2}(2a+2b). \]
503: %Since $\gcd{(a,b)},a,b \geq 1$ we really just need to e show that
504: %\[ 1\geq -ab+a+b\]
505: %on the domain in the plane $a,b\geq 1$. This is a straightforward calculus
506: %problem. \qed
507: 
508: Removing a  curve that bounds a face  does not change
509: $C_1-C_2$ or $C_1-C_3$, but $C_2+C_3-C_1$ is reduced by one. Remove
510: such curves until there are no more curves that bound faces. 
511: The remaining net will be of the form above.
512: If $\gcd{(C_1-C_2,C_1-C_3)}=2$ then the system consists of a single
513: curve. More generally, the number of components
514: is $\gcd{(a,b)}=\frac{\gcd{(C_1-C_2,C_1-C_3)}}{2}$. \qed
515: 
516: From this proof we see that the net is made up of circles that are
517: boundaries of faces along with multiple copies of a single type of circle that
518: appears in a tetrahedral net of type
519: \begin{equation} \begin{pmatrix}
520:   a & b & a+b \\ a & b & a+b 
521: \end{pmatrix}\end{equation}
522: where $a$ and $b$ are relatively prime. Alternatively, the boundary of a
523: simplex with its vertices removed is a four times punctured sphere. Any simple
524: closed curve is either boundary parallel or separates the surface into two
525: pairs of pants. The dearth of disjoint systems of simple closed curves
526: on a pair of pants causes all curves that are not triangles to be
527: parallel. The $a$ and $b$ can be understood in terms of geometric intersection
528: numbers with crosscuts.
529: Name such curves by the pair
530: $(a,b)$ where $a$ and $b$ are relatively prime and $a\leq b$.
531:  For each such
532: pair there are six or three different ways (depending on the
533: symmetries of the particular curve type) of labeling the tetrahedron
534: corresponding to the curve of type $(a,b)$.
535: We say that 
536: two $(a,b)$ curves are
537: {\em non-conflicting} if the curves are parallel in a regular neighborhood of
538: the $1$-skeleton of the tetrahedral net. 
539: 
540: \begin{prop} Euler 
541: characteristic of
542: spinal surfaces is additive when the two surfaces have the same $(a,b)$ types
543: at each vertex and those types are non-conflicting.
544: \end{prop}
545: \proof The surface that corresponds to the sum of the colorings is the
546: disjoint union of the surfaces corresponding to the two colorings. \qed
547: 
548: The type $(0,1)$ is a quad, the type
549: $(1,1)$ corresponds to an almost normal
550: surface \cite{JR}. Further types wind more and more around the 
551: tetrahedral net before closing up. 
552: 
553: 
554: 
555: %The proposition follows from 
556: 
557: %\begin{lemma}
558: % Given a tetrahedral net with curves running parallel along each edge so that
559: % the column sums of the numbers of edges are $C_{1}\geq C_{2}\geq C_{3}$,
560: %then the number of circles in in the net is greater than or equal to
561: %\[ \frac{-1}{4}(C_1-C_2)(C_1-C_3)-\frac{1}{4}(C_2+C_3)-\frac{3}{4}C_1.\]
562: %\end{lemma}
563: 
564: %\proof Notice that $\frac{\gcd{(C_1-C_2,C_1-C_3)}}{2}\geq 1$ so the number of
565: %circles is greater than or equal to $1+C_2+C_3-C_1$. If
566: %\[1+C_2+C_3-C_1 \geq
567: %\frac{-1}{4}(C_1-C_2)(C_1-C_3)-\frac{1}{4}(C_2+C_3)-
568: %\frac{3}{4}C_1,\]
569: %then we will have proved the inequality. 
570: %Putting everything on the left hand side, the inequality to prove is:
571: %\[
572: %\frac{1}{4}(C_1-C_2)(C_1-C_3)+\frac{5}{4}(C_2+C_3)-\frac{1}{4}C_1+1\geq
573: %0.\]
574: %Let $a=(C_1-C_2)/2$ and $b=(C_1-C_3)/2$. The inequality becomes,
575: %\[ (a-1)(b-1)+\frac{3}{4}(C_1+C_2+C_3).\]
576: %As long as $(a-1)(b-1)>0$ the inequality is clear. Since $b\geq a$ we
577: %only need to understand the case when $a=0$. Going back to the formula
578: %in terms of $C_i$ we see that $C_1=C_2$ and the inequality is clear.
579: 
580: \section{$6j$-symbol Details}\label{6jdet}
581: \subsection{Bounding the $6j$-symbols.}
582: 
583: We begin with a universal bound on the size of the unitary
584: $6j$--symbols in terms of their entries.
585: \begin{prop}\label{6jest}
586: Let $C_1\geq C_2\geq C_3$ be the column sums of the unitary $6j$--symbol
587: $\left\{\begin{matrix} a & b & e \\ c & d &f\end{matrix}\right\}_u$ and
588: assume that $0<t<1$.  There exists a function $K(t)>0$ such that 
589: \begin{equation}\label{fundest}
590: \left|\left\{\begin{matrix} a&b&e \\ c&d&f\end{matrix}\right\}_u\right|
591: \leq K(t)t^{\frac{1}{2}(C_1-C_2)(C_1-C_3)+C_1}.
592: \end{equation}
593: \end{prop} 
594: \proof
595: After collecting and canceling terms,
596: $\left|\left\{\begin{matrix} a&b&e \\ c&d&f\end{matrix}\right\}_u\right|$
597: is equal to 
598: \begin{equation}
599: \frac{\sqrt{\prod_v\left|[x_{v,1}]![x_{v,2}]![x_{v,3}]!\right|}}
600: {\sqrt{[\frac{a+d+e}{2}+1]![\frac{b+c+e}{2}+1]![\frac{a+b+f}{2}+1]!
601: [\frac{c+d+f}{2}+1]!}} 
602: \left|\sum_{s= m}^{M}\frac{(-1)^s[s+1]!}
603: {\prod_{i=1}^3[B_i-s]!\prod_{j=1}^4[s-A_j]!}\right|.
604: \end{equation}
605: Using (\ref{fact}) this is further equal to
606: \begin{equation}\label{first}
607: t^p(1-q)^2
608: \sqrt{\frac{\prod_v (q;q)_{x_{v,1}} (q;q)_{x_{v,2}} (q;q)_{x_{v,3}}}
609: {(q;q)_{\frac{a+d+e}{2}+1} (q;q)_{\frac{b+c+e}{2}+1}
610:   (q;q)_{\frac{a+b+f}{2}+1} (q;q)_{\frac{c+d+f}{2}+1}}} \end{equation}
611: \[
612: \left|\sum_{s=m}^M t^{p_s} \frac{1}{1-q}
613: \frac{(-1)^s (q;q)_{s+1}}
614: {\prod_{i=1}^3(q;q)_{B_i-s}\prod_{j=1}^4(q;q)_{s-A_j} }  \right|.
615: \]
616: Here 
617: \begin{equation}\label{p}
618:  p=\frac{1}{2}(-a^2-b^2-c^2-d^2-e^2-f^2 +ae+ad+ab+af +be+bc+bf +ce+cd+cf + de+df)
619: \end{equation}
620: \[+a+b+c+d+e+f
621: \]
622: and 
623: \begin{equation}
624: p_s=6s^2+ a^2+b^2+c^2+d^2+e^2+f^2\end{equation} 
625: \[+ af+ac+ae+fc+fe+ce+be+bd+bf+ed+df+ab+ad+bc+cd\]
626: \[-2s(1+2a+2b+2c+2d+2e+2f).
627: \]
628: After completing the square, $p_s$ is
629: \begin{equation}\label{ps}
630: p_s=6\left(s-\frac{a+b+c+d+e+f+\frac{1}{2}}{3}\right)^2 
631: +\frac{1}{3}(a^2+b^2+c^2+d^2+e^2+f^2)\end{equation}
632: \[- \frac{1}{3}(af+ae+ac+ad+ab+fe+fc+ce+be+bd+bf+ed+df+bc+cd)
633: -\frac{2}{3}(a+b+c+d+e+f)-\frac{1}{6}.
634: \]
635: Combining all the factors of $(1-q)$ outside and the powers of $t$ inside
636: the sum,
637: formula (\ref{first}) can be simplified to
638: \begin{equation}\label{next}
639: (1-q)
640: \sqrt{\frac{\prod_v (q;q)_{x_{v,1}} (q;q)_{x_{v,2}} (q;q)_{x_{v,3}}}
641: {(q;q)_{\frac{a+d+e}{2}+1} (q;q)_{\frac{b+c+e}{2}+1}
642:   (q;q)_{\frac{a+b+f}{2}+1} (q;q)_{\frac{c+d+f}{2}+1}}} 
643: \end{equation}
644: \[
645: \left|\sum_{s=m}^M t^{p'_s}
646: \frac{(-1)^s (q;q)_{s+1}}
647: {\prod_{i=1}^3(q;q)_{B_i-s}\prod_{j=1}^4(q;q)_{s-A_j} }  \right|.
648: \]
649: where
650: \begin{equation}\label{power}
651: p'_s=6\left(s-\frac{a+b+c+d+e+f+\frac{1}{2}}{3}\right)^2 
652: -\frac{1}{6}(a^2+b^2+c^2+d^2+e^2+f^2)\end{equation}
653: \[
654: -\frac{1}{3}(ac+fe+bd)+\frac{1}{6}(ad+ae+de+bc+be+ce+ab+af+bf+cd+cf+df)
655: \]
656: \[+\frac{1}{3}(a+b+c+d+e+f)-\frac{1}{6}.
657: \]
658: The formula (\ref{next})  is the absolute value of an alternating
659: sum from $s=m$ to $s=M$. 
660: Take the quotient whose numerator is the summand
661: at $s+1$ and whose denominator is the summand at $s$, the result is,
662: \begin{equation} (-1)t^{12s-4(a+b+c+d+e+f)+4}
663: \frac{(1-q^{s+2})\prod_{i=1}^3(1-q^{B_i-s})}{\prod_{j=1}^4(1-q^{s+1-A_j}) }.\end{equation}
664: Take the absolute value, with
665: the effect of removing the (-1). In order to see that each one of these
666: quotients is smaller than the last, take the logarithm of the
667: result, giving:
668: \begin{equation}\label{loged}
669:  (12s-4(a+b+c+d+e+f)+4)\log{(t)}+\log(1-q^{s+2})
670: +\sum_{i=1}^3 \log{(1-q^{B_i-s})}
671: -\sum_{j=1}^4 \log{(1-q^{s+1-A_j})}
672: \end{equation}
673: Apply the Taylor series for $\log{(1-x)}$ to get
674: \begin{equation}\label{log}
675:  (12s-4(a+b+c+d+e+f)+4)\log{(t)}+\sum_{n=1}^{\infty}\frac{1}{n}\left(
676: -q^{n(s+2)}-\sum_{i=1}^3q^{n(B_i-s)}+\sum_{j=1}^4 q^{n(s+1-A_j)}
677: \right).
678: \end{equation}
679: As $t<1$ the first term gets smaller as $s$ increases. We analyze 
680: the sum over $n$ in (\ref{log}) term by term. As $n$ gets larger, $q^{n(s+2)}$ gets
681: smaller so $-q^{n(s+2)}$ gets larger. However, for each $i$ and $j$, 
682: $-q^{n(B_i-s)}$
683: and $+q^{n(s+1-A_j)}$ get smaller as $s$ increases. Furthermore the powers of
684:   $q$ appearing in any $q^{n(s+1-A_j)}$ are smaller than in $-q^{n(s+2)}$
685:     which means that the amount any one of them is decreasing is greater than
686:     the amount that $-q^{n(s+2)}$ is increasing, 
687: so each term is getting smaller. Therefore
688:     the sum over all $n$ is getting smaller and the quotients are decreasing.
689: Thus, the absolute value of the sum has a
690: unique maximum. 
691: Since the sum is alternating we  conclude that the absolute value of the
692: summand is less than or
693: equal to the largest term.
694: For any $n$, using $(q;q)_n\leq 1$ in the numerator and
695: $(q;q)_n\geq (q;q)_{\infty}$ in the denominator of (\ref{next})
696: together with 
697: the fact that the power of $t$ is the
698: largest when the exponent is the smallest, the  expression in equation
699: (\ref{next}) is smaller than 
700: \begin{equation}
701: (1-q)
702: \sqrt{\frac{1}{\left((q;q)_{\infty}\right)^4}}
703: t^{\mathrm{min}}
704: \frac{1}{\left((q;q)_{\infty}\right)^7},
705: \end{equation}
706: where ${\mathrm{min}}$ is the smallest value of $p'_s$.
707: Analyzing (\ref{power}) we can see that $p'_s$ is minimal when
708:  $s=M=(C_2+C_3)/2$.
709: Substituting we see that  
710: \begin{equation}
711: p'_s\geq \frac{(C_1-C_2)(C_1-C_3)}{2}+C_1.
712: \end{equation}
713: The final estimate is
714: \begin{equation} 
715: \left|\left\{\begin{matrix} a&b&e \\ c&d&f\end{matrix}\right\}_u\right|
716: \leq
717: \frac{1-q}{\left((q;q)_{\infty}\right)^9}\cdot
718: t^{\frac{(C_1-C_2)(C_1-C_3)}{2}+C_1}. 
719: \end{equation}
720: \qed
721:                                            
722:          
723:          
724: \subsection{ Some  important limits}
725: 
726: Let $\begin{pmatrix} a & b & e \\ c & d & f \end{pmatrix}$ be an admissible
727: labeling of the edges of a tetrahedron. For any nonnegative integer $k$, the
728: labelings 
729: $\begin{pmatrix} a+2k & b+2k & e+2k \\ c+2k & d+2k & f+2k
730: \end{pmatrix}$, 
731:   $\begin{pmatrix} a+2k & b+2k & e \\ c+2k & d+2k & f+2k
732: \end{pmatrix}$, and
733: $\begin{pmatrix} a+2k & b+2k & e \\ c+2k & d+2k & f
734: \end{pmatrix}$ are admissible.
735: 
736: \begin{prop}\label{lim}
737:   Given an admissible labeling
738: $\begin{pmatrix} a & b & e \\ c & d & f \end{pmatrix}$
739:  of a tetrahedral net,
740: the sequences
741: \begin{equation}\label{seq1}
742:  t^{-4k}\left\{ \begin{matrix} a+ 2k & b + 2k & e +2k \\ c + 2k & d +2k & f
743:       +2k \end{matrix} \right\}_u, 
744: \end{equation}
745: \begin{equation}\label{seq2} 
746: t^{-4k}\left\{ \begin{matrix} a+ 2k & b + 2k & e  \\ c + 2k & d +2k & f
747:       +2k \end{matrix} \right\}_u, 
748: \end{equation}
749: and
750: \begin{equation}\label{seq3}
751:  t^{-4k}\left\{ \begin{matrix} a+ 2k & b + 2k & e  \\ c + 2k & d +2k & f
752:        \end{matrix} \right\}_u 
753: \end{equation}
754: are  convergent. 
755: \end{prop}
756: 
757: We will only prove the first limit exists, the other two are similar.
758: The proof is based on the following elementary lemma.
759: 
760: \begin{lemma}\label{altsum}
761:   Suppose that $w(k)_n$ is a sequence of sequences so that for each fixed $k$,
762:   the sequence is alternating and converges to zero, and for fixed $n$ the
763:   sequence is convergent. Suppose further that there exists $N$ so
764:   that, independent of $k$, if $n\geq N$ then $|w(k)_n|\geq |w(k)_{n+1}|$.
765:  The sequence  \begin{equation}w(k)_{\infty}=\sum_{n} w(k)_n\end{equation} 
766: (depending on $k$) is convergent.
767: \end{lemma}
768: 
769: \proof This is an application of the proof of the alternating series test.
770: \qed
771: 
772: \proof (of  Proposition \ref{lim}) 
773: Recall Formula (\ref{next}). The strand numbers increase by $1$ each time $k$
774: increases by $1$ so the $(q;q)_{x_{v,i}}$ all converge to  
775: $(q;q)_{\infty}$ as $k$ goes to infinity. Similarly, the functions in the
776: denominator inside the radical all converge to $(q;q)_{\infty}$. Hence to 
777: prove the convergence we must only understand the quantities inside the sum.
778: 
779: Let $M(k)$, $m(k)$, $A_j(k)$, $B_i(k)$ and $p'(k)_s$ be the quantities
780: in (\ref{next})
781: associated to
782: \begin{equation}\left\{ \begin{matrix} a+ 2k & b + 2k & e +2k \\ 
783: c + 2k & d +2k & f +2k \end{matrix} \right\}_u,\end{equation}
784: as in the proof of Proposition \ref{6jest}. 
785: Let $n=M(k)-s$, and let
786: \begin{equation} w(k)_n=t^{p'(k)_s-4k}
787: \frac{(-1)^s (q;q)_{s+1}}
788: {\prod_{i=1}^3(q;q)_{B_i(k)-s}\prod_{j=1}^4(q;q)_{s-A_j(k)} } \end{equation}
789: for $n\leq M(k)-m(k)$, and $w(k)_n=0$ for $n>M(k)-m(k)$.
790: 
791: As $k$ increases by $1$, the $B_i(k)$ increase by $4$ and the $A_j(k)$ only
792: increase by $3$. So, $M(k)=M(0)+4k$, $m(k)=m(0)+3k$, $B_i(k)=B_i(0)+4k$, and 
793: $A_j(k)=A_j(0)+3k$. When $n=0$, $s=M(k)$ and 
794: \begin{equation}p'(k)_{M(k)}=\frac{(C(k)_1-C(k)_2)(C(k)_1-C(k)_3)}{2}
795:   +C(k)_1,\end{equation} 
796: which increases by $4$ when $k$ increases by $1$, so 
797: $t^{-4k+p'(k)_{M(k)}}$ is a constant.
798: We see that $w(k)_0$ is convergent. A
799: similar analysis
800: shows that for fixed $n$ the sequence $w(k)_n$ is convergent. The series is
801: clearly alternating.
802: 
803: We have already seen that for fixed $k$ the sequence $|w(k)_n|$ can have at 
804: most one
805: maximum, we just need to see that there is a bound on how big $n$ is at that
806: maximum  depending only on $t$, and 
807: $\begin{pmatrix} a & b & e \\ c & d & f \end{pmatrix}$. We do this by
808: looking at $\log{|w(k)_{n}/w(k)_{n+1}|}$ and seeing when it becomes
809: nonnegative. When $n>M(0)-m(0)+k$ then $w(k)_n=0$, so the maximum of
810: $|w(k)_n|$ has already occured. Hence we only need to understand the case when
811: the quotient $w(k)_{n}/w(k)_{n+1}$ is well defined.
812: To this end we substitute into Formula (\ref{loged}) to get,
813: \begin{equation}\label{logthing}
814:  \log{|w(k)_{n}/w(k)_{n+1}|}=
815: (12(M(0)-n)-4(a+b+c+d+e+f)+4)\log{(t)}\end{equation}
816: \[+\log(1-q^{M(0)+4k-n+2})
817: -\sum_{i=1}^4\log{(1-q^{M(0)+k-n+1-A_j(0)})}+\sum_{i=1}^3
818: \log{(1-q^{(B_i(0)-M(0)+n)})}.\]
819: Choose $N$ large enough so that, 
820: \begin{equation}\label{beegtrouble}
821: (12(M(0)-N)-4(a+b+c+d+e+f)+4)\log{(t)} >\end{equation}
822: \[-\log(1-q^{M(0)+4N-N+2})-\sum_{i=1}^3\log(1-q^{M(0)-N+2}).
823: \]
824: Notice that $-\sum_{i=1}^4\log{(1-q^{M(0)+k-n+1-A_j(0)})}>0$. 
825: Inequality (\ref{beegtrouble}) guarantees that the expression
826: (\ref{logthing}) is positive when 
827: $k=N$. 
828: Increasing k in $-\log(1-q^{M(0)+4k-N+2})$
829:  makes it smaller. Thus, by the argument from the proof of
830:  Proposition \ref{6jest}, the sequence $|w_k(n)|$ is monotone decreasing for
831: $n \geq N$. 
832: %Differentiating with respect to $k$ we get
833: %\[ \sum_{j=1}^4
834: %\frac{q^{M(0)+k-n+1-A(0)_j}\log{(q)}}{1-q^{M(0)+k-n+1-A(0)_j}}-
835:  %\sum_{i=1}^3\frac{q^{B(0)_i-M(0)+n}\log{(q)}}{1-q^{B(0)_i-M(0)+n}}.\]
836: %Only the terms corresponding to $A(0)_j$ depend on $k$ and as $k$ gets larger,
837: %they get closer to zero, the other terms add up to a positive number, so there
838: %is a $K$ so that when $k\geq K$ this is
839: %positive. This means that if we choose $N$ so that all the sequences with
840: %$k\leq K$ have $|w(k)_n|$ monotone decreasing for $n\geq N$, it will be true
841: %for all $k\geq K$. 
842: 
843: We have established the criterion for convergence from  Lemma \ref{altsum}.
844: \qed
845: 
846: Let the  limit of the  sequence (\ref{seq1}) from Proposition
847: \ref{lim} be denoted by: 
848: \begin{equation}\label{sixjinf}
849: \left\{\begin{matrix} a & b & e \\ c & d & f \end{matrix}\right\}_{\infty}
850: =\lim_{k\rightarrow \infty} t^{-4k}\left\{ \begin{matrix} a+ 2k & b +
851:     2k & e +2k \\ c + 2k & d +2k & f 
852:       +2k \end{matrix} \right\}_u.
853: \end{equation}
854: Similarly, denote the limits of the sequences (\ref{seq2}) and
855: (\ref{seq3}) by 
856: \begin{equation}\label{sixjinf1}
857: \left\{\begin{matrix} a & b & e_0 \\ c & d & f \end{matrix}\right\}_{\infty}
858: =\lim_{k\rightarrow \infty}
859: t^{-4k}\left\{ \begin{matrix} a+ 2k & b + 2k & e  \\ c + 2k & d +2k & f
860:       +2k \end{matrix} \right\}_u, 
861: \end{equation}
862: and
863: \begin{equation}\label{sixjinf2}
864: \left\{\begin{matrix} a & b & e_0 \\ c & d & f_0 \end{matrix}\right\}_{\infty}
865: =\lim_{k\rightarrow \infty}
866:  t^{-4k}\left\{ \begin{matrix} a+ 2k & b + 2k & e  \\ c + 2k & d +2k & f
867:        \end{matrix} \right\}_u. 
868: \end{equation}
869: 
870: \begin{remark}
871: \begin{equation}
872:  \left\{\begin{matrix} a & b & e \\ c & d & f \end{matrix}\right\}_{\infty}
873: = 
874: (1-q)(q;q)_{\infty} \sum_{u=0}^{\infty}(-1)^{\frac{C_2+C_3}{2}+u} 
875: \end{equation}
876: \[t^{6u^2+2(2C_1-C_2-C_3+1)u+\frac{(C_1-C_2)(C_1-C_3)}{2}+C_1}
877: \frac{1}{(q;q)_u(q;q)_{u+\frac{C_1-C_2}{2}}(q;q)_{u+\frac{C_1-C_3}{2}}}\]
878: 
879: The limits (\ref{sixjinf1}) and (\ref{sixjinf2}) are zero unless $a+c=b+d$.
880: \end{remark}
881: 
882: 
883: 
884: 
885: \section{Normal and Spinal Surfaces}\label{norandspi}
886: \subsection{ Analysis of the contribution of a surface}
887: For the remainder of this paper
888: $M$ will be a compact three-manifold with
889: non-empty connected  
890: boundary. Although the method works for a more general class of manifolds,
891: this assumption simplifies the arithmetic so that the ideas behind the
892: estimates are in the foreground.
893: 
894: \begin{definition} An ideal
895: triangulation $T$ whose only normal  spheres and  tori are the link of a
896: vertex is {\bf efficient}. An ideal triangulation
897: is {\bf $0$-efficient}
898: if and only if the only embedded, normal $2$-spheres are vertex
899: linking. 
900: \end{definition}
901: The $0$-efficient triangulations were studied in \cite{JR}. In particular,
902: it is shown there that any triangulation of a closed, orientable
903: irreducible $3$-manifold can be modified to a $0$-efficient
904: triangulation, or it can be shown that the $3$-manifold is one of
905: $S^3$, ${\mathbb R}P^3$ or $L(3,1)$. It is also shown that any
906: triangulation of a compact, orientable, irreducible and boundary
907: irreducible $3$-manifold with non-empty boundary can be modified to a
908: $0$-efficient triangulation. In the announced sequel to \cite{JR}
909: authors explore the concept of {\bf $1$-efficient manifolds}. They show that
910: the triangulations of irreducible, atoroidal, closed $3$-manifolds can
911: be obtained so that in addition to being $0$-efficient,  any
912: embedded normal torus is of a very special form or the $3$-manifold is
913: $S^3$, a lens space or a small Seifert fiber space.
914: 
915: Assume that $M$ has an efficient triangulation $T$.
916: Suppose that $S$ is the spine dual to $T$.  
917: Let $F$ be a surface carried by $S$. It is induced by an admissible
918: coloring of the 
919: spine. Let  $u_f$ denote the color assigned to the  
920: face $f$.
921: At each edge $e$ the three faces sharing that edge carry colors 
922: $a_e$, $b_e$ and $c_e$.
923: At each vertex there is a corresponding coloring of a tetrahedral net,
924: $ \left(\begin{matrix}a_v & b_v & e_v \\c_v & d_v & f_v
925:   \end{matrix}\right)$. 
926: Denote the column sums at vertex $v$ by $C_{1,v}\geq C_{2,v}\geq C_{3,v}$.
927: 
928: We can form the three strand numbers at each edge: 
929: $x_{e,1}=\frac{a_e+b_e-c_e}{2}$, $x_{e,2}=\frac{c_e+b_e-a_e}{2}$  and 
930: $x_{e,3}=\frac{a_e+c_e-b_e}{2}$. 
931: These are in fact linear functionals on the
932: space of spinal surfaces.  There is an arbitrariness to the choice of
933: which function is which, so fix this choice along each edge once and for
934: all. 
935: Similarly, at each vertex
936: we can form three linear functionals $S_{1,v}$, $S_{2,v}$ and $S_{3,v}$
937: corresponding to the column sums of the tetrahedral net at the
938: vertex. 
939: \begin{definition} If $C$ is a (possibly empty) set of simple closed
940: curves on the boundary of $M$, let $\mathcal{S}(C)$ denote the set
941: of spinal surfaces with respect to an augmentation of the spine
942: corresponding to $C$. 
943: For brevity, let $\mathcal{S}= \mathcal{S}(\emptyset)$.
944: A {\bf sector} $\mathcal{F}$ is determined  by fixing the
945: order of the values of the $S_{i,v}$ at each vertex (that is, deciding
946: which of the $S_{i,v}$'s is the largest column sum, $C_{1,v}$, etc.). 
947: \end{definition}  
948: Specifying these orderings  at all vertices breaks the space of 
949: spinal surfaces into $6^{\#v}$
950: sectors. 
951: Given any infinite sequence of spinal surfaces we can find a
952: subsequence that lives in one sector, because there are only finitely
953: many sectors.
954: 
955: \begin{prop}\label{subsurf}
956:   Suppose that the spinal surface $F$ lies in the sector $\mathcal{F}$.
957: Then every connected component of $F$ lies in the same sector.
958: \end{prop}
959: 
960: \proof Recall that the intersection of a spinal surface with a tetrahedron
961: consists of triangles along with one family of disks having a particular curve
962: type $(a,b)$. The triangles contribute the same to each column of the corresponding
963: symbol so any restriction on the sector comes from the curve type. Since all
964: components of $F$ are made up of a subset of the components of the
965: intersection of $F$ with each tetrahedron, they lie in any sector that
966: $F$ lies in (and maybe some other sectors too.) \qed 
967: \begin{definition}
968: A spinal surface $F$ is {\bf $k$-peelable} if $k$ is the maximum
969: non-negative integer such that $F$ can be written as
970: $F=F'+k\cdot\partial M$.
971: Use $\mathcal{S}_k(C)$
972: to denote the set of all surfaces in $\mathcal{S}(C)$ that are $k$-peelable.
973: Similarly, use $\mathcal{F}_k(C)$ to denote the $k$-peelable surfaces
974: in the sector $\mathcal{F}(C)$.
975: \end{definition}
976: 
977: There is a one-to-one correspondence between $\mathcal{S}_0$ and
978: $\mathcal{S}_{k}$ for any $k\geq 0$ given by $u_f\rightarrow u_f+2k$
979: for every $f$. Furthermore, this correspondence preserves sectors.
980: 
981: \begin{prop}
982: A spinal  surface is in $\cS_0$ 
983: if and only if some $x_{v,i}=0$. 
984: \end{prop}
985: 
986: \proof This follows from  Proposition \ref{subtract} (on being able
987: to subtract admissible triples). \qed 
988: 
989: Consequently, a spinal surface is $k$-peelable if and only if   
990: the minimum over all strand numbers $x_{v,i}$ is equal to $k$.
991: 
992: 
993: 
994: 
995: Let $Q:\mathcal{S} \rightarrow \mathbb{Z}$
996: be the function that assigns to each surface $F$,
997: \begin{equation}\label{Q} 
998: Q(F)= \sum_f -2u_f +
999: \sum_v \frac{1}{2}(C_{1,v}-C_{2,v})(C_{1,v}-C_{3,v})+C_{1,v}.
1000: \end{equation}
1001: 
1002: \begin{prop} \label{cont}
1003: \begin{itemize}
1004: \item[{\bf(i)}] 
1005: $\displaystyle{ {-2\chi(F)}\leq Q(F)}$
1006: \item[{\bf(ii)}]  The function $Q(F)$ is super additive on any
1007: sector. That is, for any surfaces $F$, $F'$ lying in the same sector, if
1008: $C_{i,v}$ are the column sums corresponding to $F$ and $C'_{i,v}$ are the
1009: column sums corresponding to $F'$ then
1010: \begin{equation} Q(F+F')=Q(F)+Q(F')+
1011: (C_{1,v}-C_{2,v})(C'_{1,v}-C'_{3,v})/2+(C'_{1,v}-C'_{2,v})
1012: (C_{1,v}-C_{3,v})/2\end{equation}
1013: \[\geq Q(F)+Q(F')\]
1014: \item[{\bf(iii)}]  $Q(F)$ is bounded below on $\mathcal{S}_0$.
1015: \item[{\bf(iv)}]  The level sets of $Q(F)$ on $\mathcal{S}_0$ are finite.
1016: \item[{\bf(v)}]  The cardinality of the level sets of $Q$  on $\mathcal{S}_0$
1017:   grows at most polynomially in the level.
1018: \end{itemize}
1019: \end{prop}
1020: 
1021: \proof
1022: \begin{itemize}
1023: 
1024: \item[{\bf(i)}]  The Euler characteristic of the surface $F$ corresponding to the
1025: coloring $u_f$ is
1026: \begin{equation} \sum_f u_f -\sum_e \frac{a_e+b_e+c_e}{2}+\sum_v 
1027: \gcd{(C_{1,v}-C_{2,v},C_{1,v}-C_{3,v})}/2
1028: +C_{2,v}+C_{3,v}-C_{1,v} \end{equation}
1029: Because each edge has exactly two ends we can redistribute the sum to
1030: eliminate the sum over the edges. This yields,
1031: \begin{equation}\label{eq1}  
1032: \sum_f u_f+\sum_v 
1033: \gcd{(C_{1,v}-C_{2,v},C_{1,v}-C_{3,v})}/2
1034: +\frac{1}{2}C_{2,v}+\frac{1}{2}C_{3,v}-\frac{3}{2}C_{1,v}.
1035: \end{equation}
1036: Comparing (\ref{eq1}) to the right hand side of the inequality from
1037: item {\bf(i)} 
1038: we see that it is enough
1039: to show that for each vertex $v$,
1040: \begin{equation}\label{eq2}
1041: -\gcd{(C_{1,v}-C_{2,v},C_{1,v}-C_{3,v})}-C_{2,v}-C_{3,v}+3C_{1,v} \leq 
1042: \frac{1}{2}(C_{1,v}-C_{2,v})(C_{1,v}-C_{3,v})
1043: +C_{1,v}.
1044: \end{equation}
1045: In the case that $C_{1,v}-C_{2,v}=0$ this reduces to 
1046: $C_{1,v}\leq C_{1,v}$ thus the proposition is true. Assume that
1047: $C_{1,v}-C_{2,v}>0$. The triples at each vertex are admissible so
1048: $C_{1,v}-C_{2,v} \leq C_{1,v}-C_{3,v}$ are even and positive. Hence,
1049: $\gcd{(C_{1,v}-C_{2,v},C_{1,v}-C_{3,v})}\geq 2$. Substituting this in
1050: (\ref{eq2})
1051: and putting everything on the right side, the inequality is equivalent to:
1052: \begin{equation} \frac{1}{2}(C_{1,v}-C_{2,v}-2)(C_{1,v}-C_{3,v}-2)\geq 0.\end{equation}
1053: Since we are assuming $C_{1,v}-C_{2,v}\geq 2$ and
1054: $C_{1,v}-C_{3,v}\geq 2$ this is true. 
1055: 
1056: \item[{\bf(ii)}]  This is a direct computation from the formula. 
1057: 
1058: In what follows we would like to use this formula, to that end we write it
1059: more compactly as follows. Letting
1060: $F$ and $F'$ be surfaces in the same sector with
1061: $\delta_v=C_{1,v}-C_{2,v}$, $\gamma_v=C_{1,v}-C_{3,v}$ being  associated with
1062: $F$ and  $\delta_v'=C_{1,v}'-C_{2,v}'$, $\gamma_v'=C_{1,v}'-C_{3,v}'$ being associated
1063: with $F'$, we have,
1064: \begin{equation} Q(F+F')= Q(F)+Q(F')+\sum_v \frac{\delta_v\gamma_v'+\delta_v'\gamma_v}{2}.\end{equation}
1065: 
1066: 
1067: \item[{\bf(iii)}] Since there are only
1068: finitely many sectors, if  $Q$ is bounded below on each
1069: sector, then it is bounded below on $\mathcal{S}_0$.
1070: So assume we are working in a particular sector.
1071: Suppose that $Q$ is not bounded below.
1072: Starting with a surface with $Q<0$ we demonstrate the existence of
1073: another surface of a particular
1074: form with smaller $Q(F)$.
1075: We then bound $Q$ below on surfaces of that form.
1076: 
1077: Suppose that $Q(F)<0$. Decompose $F$ as a union
1078: $F_p$  of components with positive Euler characteristic 
1079: and a union $F_n$ components with negative Euler
1080: characteristic. 
1081: Since $Q(F)<0$, the surface $F_p$ is nonempty.
1082: By super-additivity we have that $Q(F)\geq
1083: Q(F_p)+Q(F_n)$. Since $Q(F_n)\geq 0$ this implies that $Q(F_p)\leq
1084: Q(F)$. Since $F_p$ is a subsurface of $F$, by Proposition
1085: \ref{subsurf} it is in the same sector. 
1086: So we can assume that we are working with a surface all of whose
1087: components are spheres.
1088: 
1089: Next assume that $F$ has $\delta_v\geq 4$ for some $v$. Our estimate that
1090: $Q(F)\geq -2\chi(F)$ tells us that if  $F$ has
1091: a single component then $Q(F)\geq -4$. Using the fact that $\delta_v\geq 4$
1092: for some vertex allows us to improve this to $Q(F)\geq -2$  Assume that
1093: $F$ is not connected. We can 
1094: then
1095: write $F=F_1+F_2$ where the $F_i$ are from $\mathcal{S}_0$, and $F_2$ is
1096: connected and has nonempty intersection with a small ball about $v$.  We use
1097: $\delta_{v,1}$, $\delta_{v,2}$ to denote the differences between the largest 
1098: column and second
1099: largest column of these two surfaces at the vertex $v$, and $\gamma_{v,1}$ and
1100: $\gamma_{v,2}$ to describe the difference between the largest column
1101: and the smallest 
1102: column. Note, 
1103: $\delta_v=\delta_{v,1}+\delta_{v,2}$ and
1104: $\gamma_v=\gamma_{v,1}+\gamma_{v,2}$. 
1105: The super-additivity formula gives
1106: \begin{equation}Q(F)=Q(F_1+F_2)=Q(F_1)+Q(F_2)+
1107: \sum_v \frac{\delta_{v,1}\gamma_{v,2}+\delta_{v,2}\gamma_{v,1}}{2}.\end{equation}
1108: Since $\delta_{v,1}+\delta_{v,2}\geq 4$ it follows that $Q(F_1)\leq Q(F)$ and 
1109: it has
1110: smaller $\delta_v$. Replace the surface $F$ with the surface $F_1$ and
1111: continue until all $\delta_v\leq 2$.
1112: 
1113: Suppose  $F$ is a surface in $\mathcal{F}_0$  with all $\delta_v\leq 2$. Since
1114: there are no normal spheres in $\mathcal{F}_0$ each sphere making up $F$ has
1115: some $\delta_v=2$. Since $\delta_v$ is additive this means that there are no 
1116: more
1117: spheres in $F$ than there are vertices in the spine. Hence $Q$ is bounded
1118: below by $-4(\text{\# vertices})$.
1119: 
1120: 
1121: \item[{\bf(iv)}] It is enough to prove that the intersection of any level set
1122:   with any sector is finite.
1123: Suppose that $F_i$ is an infinite sequence of
1124: spinal surfaces in a sector with $Q(F_i)=c$. If necessary we can pass to a
1125:   subsequence so 
1126: that  the strand numbers of the surfaces $F_i$ are
1127: monotone increasing. There are two cases. 
1128: 
1129: {\bf Case 1} If the $C_{v,1}-C_{v,2}=\delta_v$ stay
1130: bounded then we can further refine the sequence so that all these numbers are
1131: constant. As the strand numbers are monotone increasing we can subtract the
1132: first term of the sequence from every subsequent term to get a new
1133: sequence of  spinal 
1134: surfaces which are normal. The
1135: values of $Q(F_i)$ are bounded  
1136: below (by item {\bf(iii)}), hence there is an infinite sequence of surfaces
1137: with 
1138: the same Euler characteristic. Since these surfaces all  have some
1139: strand number 
1140: $0$, and the triangulation is efficient  they can be written as a sum of a 
1141: finite list of normal surfaces so
1142: that none of the surfaces has positive or zero Euler characteristic.
1143: This is a contradiction, as
1144: their Euler characteristic is increasing. 
1145: 
1146: 
1147: {\bf Case 2} If some $C_{v,1}-C_{v,2}=\delta_v$ is unbounded we refine
1148: the sequence
1149: so that the $\delta_v$ are monotone increasing and the strand numbers are 
1150: monotone
1151: increasing. Let $v$ be a vertex where  the $\delta_v$ are unbounded, and assume
1152: that the first surface in the sequence has $\delta_v>0$. If not, just start
1153: later. Subtracting the first surface from every surface in the
1154: sequence
1155: the super-additivity formula informs us that this is
1156: a sequence of surfaces in $\mathcal{S}_0$ such that $Q$ is not bounded
1157: below. This contradicts item $({\bf iii)}$.
1158: 
1159: \item[{\bf(v)}]  
1160: If $V$ is a finite dimensional free $\mathbb{Z}$-module and $v_i$ is
1161: a basis, we can define $N:V \rightarrow \mathbb{Z}$ by
1162: \begin{equation} N(\sum_i c_i v_i)=\sum_i |c_i|.\end{equation}
1163: The cardinality of the set of elements in $V$ with $N(v)\leq n$ is
1164: less than or equal to a 
1165: polynomial in $n$. Fixing a sector $\mathcal{F}$ there is a finite family of
1166: surfaces $F_i$ that generate the surfaces in $\mathcal{F}$ as an integer cone.
1167: As there are only finitely many surfaces $F$ with $Q(F)\leq0$, there is an
1168: integer $K$ so that for any $\sum_i c_i F_i$, if some $c_i\geq K$ then
1169: $Q(\sum_i c_i F_i)>0$. Let $S_j$ be the set of all surfaces $\sum_i c_i F_i$,
1170: so that some $c_i$ is between  $K$ and $2K$ and the 
1171: other $c_i$ are between $0$ and $K-1$. It is clear that all but finitely many
1172: surfaces in
1173: $\mathcal{S}$ can be written as a positive sum of these surfaces. Form
1174: a free $\mathbb{Z}$-module with basis $v_j$
1175: corresponding to the $S_j$ and define
1176: a map from the nonnegative integer sums of the $v_j$ to $\mathcal{S}$ by
1177: sending the $v_j$ to the $S_j$. This map is onto all but a finite subset
1178: of $\mathcal{F}$. Also,
1179: \begin{equation} N(\sum_i c_j v_j)\leq Q(\sum_j c_jS_j),\end{equation}
1180: so the level set $Q(S)=n$ is the image of a subset of $V$ contained inside the
1181: set $N(v)\leq n$. Therefore the level sets of $Q$ grow at most polynomially in $n$.
1182: 
1183: \end{itemize} \qed
1184: 
1185: Now suppose that $C$ is a system of simple closed curves on $\partial M$.
1186: We consider colorings of an augmentation of the spine corresponding to
1187: $C$. Let $\chi (f)$ denote the Euler characteristic of the face $f$.
1188: Note that $\chi (f)=1$ if $f$ is an open disk, and  $\chi(f)=0$ if $f$ is
1189: an annulus.
1190: 
1191: The space of surfaces corresponding to admissible colorings of the augmented
1192: spine is much like the space of spinal surfaces, except you can't add two
1193: augmented colorings. However, you can add an augmented coloring and any
1194: coloring of the original spine. We can divide the space of colorings of the
1195: augmented spine into sectors just like we did for spinal surfaces, and we can
1196: define $k$-peelable. Let $\mathcal{F}(C)$ be the surfaces in a
1197: sector coming from 
1198: colorings of an augmentation of the spine,
1199: and denote by $\mathcal{F}$ the corresponding sector in space of
1200: spinal surfaces associated to the original spine. Use  
1201: $\mathcal{F}(C)_k$ to denote the  $k$-peelable surfaces in that sector.
1202: Define $Q$ from the space of surfaces
1203: corresponding to admissible colorings of the augmented spine to the counting
1204: numbers by,
1205: \begin{equation}\label{newQ} 
1206: Q(F)= \sum_f -2 \chi (f) u_f +
1207: \sum_v \frac{1}{2}(C_{1,v}-C_{2,v})(C_{1,v}-C_{3,v})+C_{1,v}.
1208: \end{equation}
1209: 
1210: 
1211: \begin{prop}\label{aug}
1212: \begin{itemize}
1213: \item[{\bf(i)}]  $ {-2x_F}\leq Q(F)$
1214: \item[{\bf(ii)}]  The function $Q(F)$ is super additive on sectors.
1215:  If
1216: $C_{i,v}$ are the column sums corresponding to $F \in \mathcal{F}(C)$
1217:  and $C'_{i,v}$ are the 
1218: column sums corresponding to $F'\in \mathcal{F}$ then
1219: \begin{equation} Q(F+F')=Q(F)+Q(F')+
1220: (C_{1,v}-C_{2,v})(C'_{1,v}-C'_{3,v})/2+(C'_{1,v}-C'_{2,v})(C_{1,v}-C_{3,v})/2\end{equation}  
1221: \[\geq Q(F)+Q(F').\]
1222: \item[{\bf(iii)}]  $Q(F)$ is bounded below on $\mathcal{S}(C)_k$.
1223: \item[{\bf(iv)}]  The level sets of $Q(F)$ on $\mathcal{S}(C)_k$ are finite.
1224: \item[{\bf(v)}]  The cardinality of the level sets of $Q$  on $\mathcal{S}(C)_k$ grows
1225: at most polynomially in the level.
1226: \end{itemize}
1227: \end{prop}
1228: 
1229: \proof The proof is an extension of the proof of  Proposition \ref{cont}. The
1230: first two parts follow directly from the formula for $Q$. 
1231: 
1232: The third part we
1233: argue  as follows. First get the estimate on $\mathcal{S}(C)_0$
1234: by working in sectors. 
1235: Given a surface $F \in \mathcal{F}(C)_0$ it can be written as a sum
1236: of a surface $F_1$ such that each of its components has nonempty boundary and
1237: a surface $F_2$ each
1238: component of which is closed. From Proposition \ref{cont} we have a lower
1239: bound for $Q(F_2)$, from inequality {\bf(i)}  we
1240: can bound $Q(F_1)$ below 
1241: by $-2$ times the number of components in $C$. By super-additivity we have
1242: bounded $Q$ from below on $\mathcal{F}(C)_0$. 
1243: 
1244: To bound $Q$ below on $\mathcal{S}(C)_k$ use the one-to-one correspondence
1245: between surfaces in $\mathcal{S}(C)_k$ and surfaces in $\mathcal{S}(C)_0$
1246: obtained by adding $k$ copies of $\partial M$. Once again we bound the value
1247: of $Q$ on $k$ parallel copies of the boundary using the inequality from
1248: item {\bf(i)} 
1249: and then use the bound on $\mathcal{S}(C)_0$ and super-additivity
1250: on sectors.
1251: 
1252: The proofs of items {\bf(iv)}  and {\bf(v)}  are completely analogous
1253: to the proofs in  Proposition \ref{cont}.
1254: \qed 
1255: 
1256: 
1257: 
1258: \subsection{Summing Over  $k$-peelable Surfaces}
1259:  Let $C$ be a system of simple closed curves in $\partial M$,  let
1260: $S$ be a spine that is dual to an efficient triangulation of $M$
1261: and let $S(C)$ be an augmentation of $S$ with respect to $C$. 
1262: Given a coloring  $F$  of the augmented spine $S(C)$  let the $u_f$,
1263: $a_e$, $b_e$, $c_e$, $a_v$, $b_v$, $c_v$, $d_v$, $e_v$, $f_v$ and $\chi(f)$ be
1264: defined as before. 
1265: Also, let $\chi(e)=1$ if the edge $e$ has some vertex in its closure and 
1266: let $\chi(e)=0$ otherwise ($\chi(e)$ is the Euler characteristic of the
1267: edge $e$).
1268: \begin{definition}
1269: The contribution of $F$ is defined to be
1270: \begin{equation} E(F)=\frac{\prod_f \Delta_{u_f}^{\chi(f)} \prod_v \Tet a_v & b_v & e_v \\
1271: c_v & d_v & f_v \end{pmatrix}}{\prod_e \theta(a_e,b_e,c_e)^{\chi(e)}}.\end{equation}
1272: \end{definition}
1273: Notice that faces 
1274: and edges of the spine contribute to $E(F)$ unless they are annular
1275: or belong to the simple closed curves on the boundary respectively. 
1276: Each vertex is an endpoint of four edges and each edge
1277: that counts in the contribution of a surface has two ends. We can thus
1278: collect the tetrahedral coefficient at each vertex with the thetas to
1279: reparse this product as
1280: \begin{equation}\label{reparse} 
1281: E(F)= \prod_f \Delta_{u_f}^{\chi(f)}\prod_v \left\{ 
1282:     \begin{matrix}
1283:      a_v & b_v & e_v \\
1284: c_v & d_v & f_v 
1285:     \end{matrix}\right\}_u.
1286: \end{equation} 
1287: 
1288: There is a map $S(C)_0 \rightarrow S(C)_k$ that adds $k$ copies of the
1289: boundary of $M$ (as a union of triangles near the vertex). This map is one to one
1290: and onto. If the largest color corresponding to $F$ is $N$ then the largest
1291: color corresponding to $F+k\partial M$ is $N+2k$.
1292: We define 
1293: \begin{equation}\label{Ek}
1294: E_k(F)=E(F+k\partial M).
1295: \end{equation}
1296: Since
1297: $\chi(M)=\#f-\#v$,
1298: \begin{equation}\label{kpeel}Q(F+k\partial
1299: M)=Q(F)-4k\chi(M).\end{equation}
1300: Using results of Proposition \ref{lim} about limits of $6j$ symbols we have,
1301: \begin{prop}\label{Es}
1302: For every surface $F \in \mathcal{S}(C)_0$, the limit
1303: \begin{equation}\label{Einf}
1304: \lim_{k\rightarrow\infty}t^{4k\chi(M)}E_k(F)
1305: \end{equation}
1306:  exists. When $C=\emptyset$, it is equal to
1307: \begin{equation}\label{limeinf}
1308: E_{\infty}(F)= 
1309: \prod_f (-1)^{u_f}
1310: \frac{t^{-2u_f}}{1-q}
1311: \prod_{v}\left\{  \begin{matrix}
1312:      a_v & b_v & e_v \\
1313: c_v & d_v & f_v 
1314:     \end{matrix}\right\}_{\infty}. 
1315: \end{equation}
1316: \end{prop}
1317: 
1318: \proof
1319: 
1320: Assume first that $C=\emptyset$, thus $\chi(f)=1$ for all $f$.
1321: Given a surface $F \in \mathcal{S}(C)_0$  and  $k>0$, 
1322: use (\ref{reparse}) together with (\ref{Delta})
1323: to express  
1324: \begin{equation}\label{firste}
1325: E_k(F)=\prod_{f} (-1)^{-u_f-2k}t^{-2u_f-4k} \frac{1-q^{u_f+2k}}{1-q}  
1326: \prod_v t^{4k}t^{-4k}
1327: \left\{ 
1328:     \begin{matrix}
1329:      a_v+2k & b_v+2k & e_v+2k \\
1330: c_v+2k & d_v+2k & f_v+2k 
1331:     \end{matrix}\right\}_u.
1332: \end{equation}
1333: Since $\chi(M)=\#f-\#v$, equation (\ref{firste}) can be rewritten as
1334: \begin{equation}\label{seconde}
1335: t^{-4k\chi(M)}\prod_f (-1)^{u_f}t^{-2u_f}\frac{1-q^{u_f+2k}}{1-q} 
1336: \prod_v t^{-4k} 
1337: \left\{  \begin{matrix}
1338:      a_v+2k & b_v+2k & e_v+2k \\
1339: c_v+2k & d_v+2k & f_v+2k 
1340:     \end{matrix}\right\}_u.
1341: \end{equation}
1342: By Proposition \ref{lim}, along with the fact that
1343: $\lim_{k\rightarrow\infty}\frac{1-q^{u_f+2k}}{1-q}=\frac{1}{1-q}$,  
1344: limit  (\ref{Einf}) exists and is
1345: given by the formula (\ref{limeinf}).
1346: 
1347: In the case when $C\neq\emptyset$ the argument is similar. The product
1348: in (\ref{firste}) must be taken over all faces $f$ with $\chi(f)\neq
1349: 0$ and for some of the vertices $v$ we need to consider the limit of 
1350: sequences (\ref{seq2}) or (\ref{seq3}) instead of the sequence
1351: (\ref{seq1}) as in equation (\ref{seconde}).
1352:    
1353: 
1354: 
1355: \qed
1356: 
1357: Let $\mathcal{S}(C)_k^N$ be the subset of $\mathcal{S}(C)_k$
1358:  where the largest color $u_f$ is less than or equal to $N$ and let
1359: $\mathcal{S}(C)_k^{T(N)}$ be the subset of $\mathcal{S}(C)_k$
1360: where the largest $u_f$ is greater than $N$, the {\em tail} of the set.
1361: Clearly, 
1362: \begin{equation}\mathcal{S}(C)_k=\mathcal{S}(C)_k^N\cup \mathcal{S}(C)_k^{T(N)}.\end{equation}
1363: 
1364: \begin{lemma}\label{tail}
1365:   For every $\epsilon >0$ there is $N$ so that for all $k$,
1366: \begin{equation} 
1367: \sum_{F \in \mathcal{S}(C)_k^{T(N+2k)}} |E(F)|< t^{-4k \chi(M)} \epsilon,
1368: \end{equation}
1369: where $\chi(M)$ is the Euler characteristic of the manifold $M$.
1370: 
1371: Moreover, 
1372: for every $i\geq 0$, the limit 
1373: \begin{equation}\label{Fki}
1374: \lim_{k\rightarrow\infty}t^{4k\chi(M)}\sum_{F \in
1375:   \mathcal{S}(C)_k^{k+i}} |E(F)|
1376: \end{equation}  
1377: exists.
1378: 
1379: \end{lemma}
1380: 
1381: 
1382: \proof 
1383: Using (\ref{reparse})
1384:  along with the estimate from Proposition
1385: \ref{6jest}, we see that, 
1386: \begin{equation} |E(F)| \leq D(t,M,C) t^{Q(F)},\end{equation}
1387: where $D(t,M,C)$ is a number that only depends on $t$, the manifold $M$ and
1388: the augmentation of the spine corresponding to $C$. From Proposition \ref{aug}
1389: the function $Q(F)$ is bounded below by some $Q_0 \in \mathbb{Z}$,
1390: and has finite level sets, so that the
1391: level set where $Q$ takes on the value $n$ has its cardinality bounded above
1392: by a polynomial $p(n)$. Comparing with 
1393: \begin{equation} \sum_{n\geq Q_0} p(n)t^n,\end{equation}
1394: the series 
1395: \begin{equation}\sum_{F \in \mathcal{S}(C)_0} |E(F)|\end{equation}
1396: is absolutely summable. This means that for each $\epsilon>0$ there is
1397: $N$ so that 
1398: \begin{equation} \sum_{F \in \mathcal{S}(C)_0^{T(N)}} D(t,M,C) t^{Q(F)}<  \epsilon.\end{equation}
1399: Using equation (\ref{kpeel})
1400: we have
1401: \begin{equation} \sum_{F \in \mathcal{S}(C)_k^{T(N+2k)}} D(t,M,C) t^{Q(F)}<  
1402: t^{-4k\chi(M)}\epsilon.\end{equation} 
1403: 
1404: Combining the above argument with Proposition \ref{Es} yields the
1405: existence of the limit (\ref{Fki}).
1406: \qed
1407: 
1408: 
1409: \begin{remark}\label{restatedlemma}
1410:   The first part of this lemma can be restated as follows: for every 
1411: $\epsilon>0$ there exists $N$ so that independent of $k$,
1412: \begin{equation} \sum_{F \in \mathcal{S}(C)_0^{T(N)}} 
1413: t^{4k\chi(\partial M)} |E_k(F)|<\epsilon.\end{equation}
1414: 
1415: \end{remark}
1416: 
1417: 
1418: 
1419: 
1420: 
1421: 
1422: \begin{prop}\label{THEZ}
1423: Let 
1424: \begin{equation}
1425: Z_k(M)=\sum_{F \in \mathcal{S}(C)_k} E(F)=
1426: \sum_{F \in \mathcal{S}(C)_0} E_k(F),
1427: \end{equation}
1428: and
1429: \begin{equation}
1430: |Z|_k(M)=\sum_{F \in \mathcal{S}(C)_k} |E(F)|=
1431: \sum_{F \in \mathcal{S}(C)_0} |E_k(F)|=.
1432: \end{equation}
1433: For each $k$, $Z_k(M)$ and $|Z|_k(M)$ are well defined.
1434: Moreover, the limits 
1435: $Z_{\infty}(M)=\lim_{k\rightarrow \infty} t^{4k\chi(M)}Z_k(M)$
1436: and $|Z|_{\infty}(M)=\lim_{k\rightarrow \infty} t^{4k\chi(M)}|Z|_k(M)$
1437: exist.
1438: \end{prop}
1439: 
1440: \proof
1441: The well defined part of the proposition follows directly from Lemma
1442: \ref{tail}.
1443: 
1444: In order to prove convergence of $Z_k(M)$,
1445: choose $\epsilon >0$. There exists $N$ so that for all $k$
1446: \begin{equation} \sum_{F \in \mathcal{S}(C)_k^{T(N+2k)}} D(t,M,C) t^{Q(F)}<  
1447: t^{-4k\chi(M)}\epsilon/4.\end{equation}
1448: By Proposition \ref{Es} there is a $K$ so that if $k_1,k_2>K$ then
1449: \begin{equation} | t^{4k_1\chi(M)}\sum_{F \in\mathcal{S}(C)_{k_1}}^NE(F)- 
1450: t^{4k_2\chi(M)}\sum_{F \in \mathcal{S}(C)_{k_2}}^NE(F)|<\epsilon/2.\end{equation}
1451: This means that
1452: \begin{equation} |t^{4k_1\chi(M)}Z_{k_1}(M)-t^{4k_2\chi(M)}Z_{k_2}(M)|\leq \end{equation}
1453: \[ | t^{4k_1\chi(M)}\sum_{F \in \mathcal{S}(C)_{k_1}}^NE(F)- 
1454: t^{4k_2\chi(M)}\sum_{F \in\mathcal{S}(C)_{k_2}}^NE(F)|+\]
1455: \[|t^{4k_1\chi(M)}\sum_{F \in \mathcal{S}(C)_{k_1}^{T(N+2k_1)}}E(F)|+
1456: |t^{4k_2\chi(M)}\sum_{F \in \mathcal{S}(C)_{k_2}^{T(N+2k_2)}}E(F)|\leq\]
1457: \[ \epsilon/2 +\epsilon/4+\epsilon/4.\] As the sequence is Cauchy it
1458: converges. The same proof works for $|Z|_{\infty}$. \qed
1459: 
1460: \section{The Invariant Sums}
1461: In the section we analyze the sum of contributions of all spinal
1462: surfaces in the three-manifold $M$ with an efficient triangulation.
1463: 
1464: Given any integer $r\geq 3$,  all the
1465: special functions, $\Delta_n$, $\theta(a,b,c)$,
1466: $\text{Tet}\begin{pmatrix} a & b & e \\ c & d &
1467: f\end{pmatrix}$, are well defined for $t=e^{\frac{\pi i}{2r}}$
1468: whenever $a,b,c,d,e,f\leq r-1$ and the 
1469: condition $a+b+c\leq 2r-4$ is added to the definition of
1470: admissibility. Given a system $C$ of disjoint simple closed curves in
1471: $\partial M$ and  an augmentation $S(C)$ of the spine dual to the
1472: triangulation of $M$,
1473: the (finite) sum over all $r$-admissible colorings of the faces
1474: of $S(C)$,
1475: \begin{equation}\label{turvir}
1476: \sum_{\text{$r$-admissible colorings of $S(C)$}}
1477:   \frac{\prod_ f\Delta_{u_f}^{\chi(f)} \prod_v \Tet a_v & b_v & e_v \\
1478:     c_v & d_v & f_v 
1479: \end{pmatrix} } {\prod_e \theta(a_e,b_e,c_e)^{\chi(e)}},
1480: \end{equation}
1481: is a coefficient (corresponding to $C$) of a vector-valued invariant
1482: associated to $M$ by the topological quantum field theory underlying
1483: the Turaev-Viro invariant of $M$ at level $r$. Our idea is the extend
1484: the invariant away from the roots of unity. The first major step is to
1485: analyze the convergence of the infinite sums like (\ref{turvir}),
1486: where $t=e^{\frac{\pi i}{2r}}$ is replaced by any $0<t<1$ and the
1487: colorings are admissible.
1488: 
1489: 
1490: 
1491: 
1492: \begin{theorem}
1493: Let $S(C)^N$ denote the set of admissible colorings of $S(C)$
1494: with all $u_f\leq N$.
1495: \begin{itemize}
1496: \item[{\bf(i)}]   If the Euler characteristic of $M$ is negative then
1497: \begin{equation}\label{bigen}
1498:  \sum_{\text{admissible colorings $u_f$ of $S(C)$}}
1499:   \frac{\prod_ f\Delta_{u_f}^{\chi(f)} \prod_v \Tet a_v & b_v & e_v \\
1500:     c_v & d_v & f_v 
1501: \end{pmatrix} } {\prod_e \theta(a_e,b_e,c_e)^{\chi(e)}}
1502: \end{equation}
1503: converges absolutely.
1504: \item[{\bf(ii)}] If $\chi(M)=0$ then
1505: \begin{equation}\label{knotcomp} 
1506: \lim_{N\rightarrow \infty}\frac{1}{N}
1507: \sum_{S(C)^N}
1508:   \frac{\prod_{f}\Delta_{u_f}^{\chi(f)} \prod_v \Tet a_v & b_v & e_v
1509:     \\ c_v & d_v & f_v 
1510: \end{pmatrix}}{\prod_e \theta(a_e,b_e,c_e)^{\chi(e)}}
1511: \end{equation}
1512: exists and is equal to $Z_{\infty}(M)=\sum_{F\in S_0(C)}E_{\infty}(F)$
1513: which converges  absolutely. 
1514: \item[{\bf(iii)}] 
1515: If $\chi(M)=1$ then
1516: \begin{equation}\label{sphere} 
1517: \lim_{N\rightarrow \infty} t^{8N}
1518: \sum_{S(C)^{2N}}
1519:   \frac{\prod_{f}\Delta_{u_f}^{\chi(f)} \prod_v \Tet a_v & b_v & e_v
1520:     \\ c_v & d_v & f_v 
1521: \end{pmatrix}}{\prod_e \theta(a_e,b_e,c_e)^{\chi(e)}}
1522: \end{equation}
1523: exists. 
1524: Given a spinal surface $F$, let $m(F)$ denote the least even number
1525: greater than or equal to the  maximal
1526: color corresponding to $F$.
1527: The limit (\ref{sphere}) is equal to the sum of the absolutely
1528: convergent series: 
1529: \begin{equation}\label{spherelim}
1530: \frac{1}{1-q} 
1531: \sum_{F\in S_0(C)}
1532: t^{4m(F)}
1533: E_{\infty}(F).
1534: \end{equation} 
1535: \end{itemize}
1536: \end{theorem}
1537: 
1538: \proof 
1539: \begin{itemize}
1540: \item[{\bf(i)}]
1541: We need to show that the sequence of partial sums of the absolute
1542: values of the  series
1543: (\ref{bigen}) converges, that is,  
1544: \begin{equation} \lim_{N\rightarrow \infty}
1545: \sum_{S(C)^N}
1546:  \left| \frac{\prod_{f}\Delta_{u_f}^{\chi(f)} \prod_v \Tet a_v & b_v & e_v
1547:     \\ c_v & d_v & f_v 
1548: \end{pmatrix}}{\prod_e \theta(a_e,b_e,c_e)^{\chi(e)}}\right|\end{equation}
1549: exists.
1550: Notice that 
1551: \begin{equation}\sum_{S(C)^N}
1552: \left|  \frac{\prod_{f}\Delta_{u_f}^{\chi(f)} \prod_v \Tet a_v & b_v & e_v
1553:     \\ c_v & d_v & f_v 
1554: \end{pmatrix}}{\prod_e \theta(a_e,b_e,c_e)^{\chi(e)}}\right|=\end{equation}
1555: \[\sum_k 
1556: \sum_{F \in \mathcal{S}(C)_k^N} |E(F)|<
1557: \sum_k |Z|_k(M).\]
1558: Proposition \ref{THEZ} implies that the series $\sum_k|Z|_k(M)$
1559: converges by comparison with the series $\sum_kt^{-4k\chi(M)}$.
1560: 
1561: \item[{\bf(ii)}]
1562: First, regroup the finite sum in (\ref{knotcomp}) according to
1563: $k$-peelable surfaces. That 
1564: is, use the fact that $S(C)^N$ is a disjoint union of subsets
1565: $S(C)_k^N$ with $k=0,\dots ,N$ (since $\mathcal{S}(C)_k^N$ is empty for
1566: $k>N$). Thus,
1567: \begin{equation}\label{regroup}
1568: \frac{1}{N}
1569: \sum_{S(C)^N}
1570:   \frac{\prod_{f}\Delta_{u_f}^{\chi(f)} \prod_v \Tet a_v & b_v & e_v
1571:     \\ c_v & d_v & f_v 
1572: \end{pmatrix}}{\prod_e \theta(a_e,b_e,c_e)^{\chi(e)}}=
1573: \frac{1}{N}
1574: \sum_{k=0}^{N}\sum_{F \in \mathcal{S}(C)_k^N} E(F).
1575: \end{equation}
1576: %Given $\epsilon>0$ we need to find $n$ so that for all $N>n$
1577: %\[\left|Z_{\infty}(M)-\frac{1}{N}\sum_{k=0}^N
1578: %\sum_{F \in \mathcal{S}(C)_k^N} E(F)\right|<\epsilon
1579: %\]
1580: By Proposition \ref{THEZ} we can find $K$ so that for all $k>K$ we
1581: have 
1582: \begin{equation}\label{zbound}
1583: |Z_k(M)-Z_{\infty}(M)|<\frac{\epsilon}{4}.\end{equation}
1584: By Lemma \ref{tail} there exists $n_1$ so that for all $k$
1585: \begin{equation}\label{kbound}
1586: |Z_k(M)-\sum_{F \in \mathcal{S}(C)_k^{n_1+k}}
1587: E(F)|<\frac{\epsilon}{4}.\end{equation}
1588: Combining these, we get that for all $k>K$, all $n_0\geq n_1$
1589: 
1590: \begin{equation}|Z_{\infty}(M)-\sum_{F\in\mathcal{S}(C)_k^{k+n_0}}E(F)|<\frac{\epsilon}{2}.\end{equation}
1591: Therefore, each of the $N-K-n_1-1$ terms of the sum
1592: \begin{equation}\sum_{k=K+1}^{N-n_1}\sum_{F \in \mathcal{S}(C)_k^N} E(F)\end{equation}
1593: is at most $\frac{\epsilon}{2}$ away from $Z_{\infty}(M)$. Since
1594: $ \lim_{N\rightarrow \infty}\frac{N-K-n_1-1}{N}=1$ to finish the proof
1595: it suffices to show that the first $K+1$ terms and the last $n_1$ terms inside the
1596: outer sum on the right hand side of (\ref{regroup}) are
1597: bounded regardless of the value of $N$.
1598: For the first $K+1$ terms notice that  by (\ref{kbound})
1599: \begin{equation}|\sum_{k=0}^K\sum_{F \in \mathcal{S}(C)_k^N} E(F)|<K(\frac{\epsilon}{4}+B),\end{equation}
1600: where $B=\text{max}(|Z_0(M)|,|Z_1(M)|,\dots |Z_K(M)|)$.
1601: The fact that the last $n_1$ inner sums 
1602: \begin{equation}|\sum_{k=N-n_1}^N\sum_{F \in \mathcal{S}(C)_k^N} E(F)|\end{equation}
1603: are bounded
1604: regardless of $N$ follows from the fact that for every $i$
1605: the limit 
1606: \begin{equation}\lim_{k\rightarrow\infty}\sum_{F \in \mathcal{S}(C)_k^{k+i}} E(F)\end{equation}
1607: exists (see Lemma \ref{tail}).
1608: 
1609: \item[{\bf(iii)}]
1610: Absolute convergence of the sum (\ref{spherelim}) follows from the existence of the
1611: universal bound on $|E_{\infty}(F)|$ for $F \in S_0(C)$. Since
1612: $\lim_{k \rightarrow \infty}t^{4k}E_k(F)=E_{\infty}(F)$, this in turn
1613: follows from a universal bound on $|t^{4k}E_k(F)|$ 
1614: for $F \in S_0(C)$. 
1615: By letting
1616: $\epsilon=\frac{1}{2}$ in Remark
1617: \ref{restatedlemma} we see that except for finitely many surfaces $F \in S_0(C)$,
1618: $t^{4k}E_k(F)<\frac{1}{2}$. Since each of the sequences $t^{4k}E_k(F)$
1619: is convergent for the remaining surfaces, the quantities $|t^{4k}E_k(F)|$ 
1620:  are universally bounded for all surfaces $F \in S_0(C)$.
1621: 
1622: 
1623: Our goal is to show that the sequence
1624: \begin{equation}\label{finsum} 
1625:  t^{8N}\sum_{F \in
1626:     \mathcal{S}(C)^{2N}} E(F)
1627: \end{equation} 
1628: converges to  the sum (\ref{spherelim}). The first step is
1629: to rewrite the 
1630: finite sum in (\ref{finsum}) so that it is a sum over $0$-peelable
1631: surfaces. We 
1632: get,
1633: \begin{equation} 
1634: \sum_{F \in \cS_0(C)}\sum_{k=0}^{2N-m(F)} t^{8N}E_k(F).
1635: \end{equation}
1636: The largest part of this sum is at the end, so we change variables to put
1637: the largest part at the beginning. Let $i=2N-m(F)-k$. Substitution,
1638: along with splitting off an appropriate power of $t$, yields:
1639: \begin{equation}\label{refinsum} 
1640: \sum_{F \in \cS_0(C)}\sum_{i=0}^{2N-m(F)}t^{4m(F)+4i}
1641:  t^{8N-4m(F)-4i}E_{2N-m(F)-i}(F).\end{equation}
1642: From Remark \ref{restatedlemma} there exists $K_0$ so that, for all
1643: $i\geq K_0$,
1644: \begin{equation}  
1645: \sum_{F \in \cS_0(C)^{T(K_0)}}t^{4i\chi(M)}E_i(F)<\frac{\epsilon (1-q)}{4}, 
1646: \end{equation} thus
1647: \begin{equation}  \sum_{F \in \cS_0(C)^{T(K_0)}}\frac{t^{4m(F)}}{1-q}
1648: E_i(F)<\frac{\epsilon
1649:   }{4}.\end{equation}
1650: Estimating based on summing the geometric series $\sum_i t^{4i}=\frac{1}{1-q}$
1651: we can truncate the sum (\ref{refinsum})  using any $K\geq K_0$  as follows and remain within 
1652: $\epsilon/4$ of the
1653: original sum.
1654: \begin{equation}\label{firstrunc} 
1655: \sum_{F \in \cS_0(C)^{K}}\sum_{i=0}^{2N-m(F)}t^{4m(F)+4i}
1656:  t^{8N-4m(F)-4i}E_{2N-m(F)-i}(F).\end{equation}
1657: Since by Proposition \ref{aug} the function $Q(F)$ is bounded below on
1658: $\cS_0(C)$ there exists $B$
1659: so that for all $F$, $N$ and $i$
1660: \begin{equation} t^{8N-4m(F)-4i}E_{2N-m(F)-i}(F)<B.\end{equation}
1661: From the elementary theory of the geometric series there exists
1662: $I$ so that for all $F\in \cS_0(C)^{K}$,
1663: \begin{equation} 
1664: \sum_{i\geq I}^{2N-m(f)} t^{4m(F)+4i} B <\epsilon/4.\end{equation}
1665: This means we can truncate the  sum (\ref{firstrunc}) again as follows and
1666: remain within $\epsilon/4$ of 
1667: the original sum:
1668: \begin{equation}\label{secontruc}
1669:  \sum_{F \in \cS_0(C)^{K}}\sum_{i=0}^{I}t^{4m(F)+4i}
1670:  t^{8N-4m(F)-4i}E_{2N-m(F)-i}(F).\end{equation}
1671: Using the fact that for any $F$, and for any fixed $i$,
1672: \begin{equation} 
1673: \lim_{N\rightarrow \infty}
1674: t^{8N-4m(F)-4i}E_{2N-m(F)-i}(F)=E_{\infty}(F),
1675: \end{equation}
1676: together with the fact that the number of terms of the sum
1677: (\ref{secontruc})  is bounded
1678: independent of 
1679: $N$, we can choose $N$ so large that the sum (\ref{secontruc}) is
1680: within $\epsilon/4$ of
1681: \begin{equation}\label{penultimate} 
1682: \sum_{F \in \cS_0(C)^{K}}\sum_{i=0}^{I}t^{4m(F)+4i}
1683:  E_{\infty}(F),\end{equation}
1684: leaving us within $\frac{3\epsilon}{4}$ of the original sum
1685: (\ref{finsum}). Using the 
1686: absolute convergence of $\sum_{F \in \cS_0(C)} t^{4m(F)}E_{\infty}(F)$,
1687: and the fact that the bound $B$ is still valid for $E_{\infty}(F)$,
1688: we can choose $I$ large enough to make this last sum (\ref{penultimate}) 
1689: within $\epsilon/4$ of 
1690: \begin{equation} 
1691: \sum_{F \in \cS_0(C)}\sum_{i=0}^{\infty}t^{4m(F)+4i}
1692:  E_{\infty}(F).\end{equation}
1693: Summing the geometric series yields the final result. \qed
1694: 
1695: 
1696: \end{itemize}
1697: 
1698: \begin{thebibliography}{9999}
1699: \bibitem{BP} R. Benedetti, C. Petronio, {\em Branched Standard Spines of
1700:     Three-Manifolds}, Lecture Notes in Math 1653, Springer-Verlag,
1701:   Berlin-Heidelberg-New York, 1997.
1702: 
1703: \bibitem{BFK}  D. Bullock, C. Frohman, J. Kania-Bartoszynska, {\em
1704:     Understanding the Kauffman Bracket Skein Module}, J. Knot Theory
1705: Ramifications 8 (1999), no. 3, 265--277.
1706: 
1707: \bibitem{nefeli} D. Bullock, C. Frohman, J. Kania-Bartoszynska, {\em
1708:     The Yang-Mills measure in the Kauffman bracket skein module},
1709:   Comment. Math. Helv. {\bf 78} (2003) 1-17
1710: 
1711: \bibitem{ob} C. Frohman, J. Kania-Bartoszynska, {\em A Quantum
1712:     Obstruction to Embedding},   
1713: Math. Proc. Camb. Phil. Soc.  {\bf 131}, (2001), 279-293.
1714: 
1715: 
1716: \bibitem{GR} G. Gasper, M. Rahman, {\em Basic Hypergeometric Series},
1717:   Cambridge University Press, 1990.
1718: 
1719: \bibitem{JR} W. Jaco, H. Rubinstein, {\em 0-Efficient Triangulations of
1720:     Three-Manifolds}, xxx.lanl.gov/math.GT/0207158.
1721: 
1722: \bibitem{KL} L. H. Kauffman and S. Lins, {\em Temperley-Lieb
1723: recoupling theory and invariants of $3$-manifolds}, Ann.\ of Math.\
1724: Studies {\bf 143}, Princeton University Press, 1994.
1725: 
1726: \bibitem{Ma} S.V. Matveev, {\em Transformations of Special Spines and the
1727:     Zeeman Conjecture}, Math. USSR-Izv. {\bf 31} (1988), 423-434.
1728: 
1729: \bibitem{MV} G. Masbaum, P. Vogel, {\em 3-valent graphs and the Kauffman 
1730: bracket.},
1731: Pacific J. Math. {\bf 164} (1994) 361-381.
1732: 
1733: 
1734: \bibitem{Pi} R. Pergallini, {\em Standard Moves for Standard Polyhedra and
1735:     Spines}, Rendicotti Circ. Mat. Palermo {\bf 37} suppl. 18 (1988), 391-414.
1736: 
1737: \bibitem{Ro} J. Roberts, {\em Classical $6j$-Symbols and the
1738:     Tetrahedron}, Geometry and Topology, {\bf 3} (1999), pages 21--66.
1739:   
1740: \bibitem{TV} V.~G. Turaev and O.~Y. Viro,
1741: {\it State sum invariants of 3-manifolds and quantum 6j-symbols}\/,
1742: Topology, {\bf 31} (1992), 866--902.
1743: 
1744: \end{thebibliography}
1745: \end{document}
1746: 
1747: 
1748: