1: \documentclass[12pt]{article}
2: \setlength{\textheight}{23cm}
3: \pagestyle{plain}
4: \usepackage[dvips]{graphicx}
5: \usepackage{amsmath,amsthm,amsfonts,amssymb,amscd,epic,eepic}
6:
7:
8: \newtheorem{thm}{Theorem}[section]
9: \newtheorem{cor}[thm]{Corollary}
10: \newtheorem{defn}[thm]{Definition}
11: \newtheorem{exm}[thm]{Example}
12: \newtheorem{lem}[thm]{Lemma}
13: \newtheorem{prop}[thm]{Proposition}
14: \newtheorem{ack}[thm]{Acknowledgments}
15: \newtheorem{property}[thm]{Property}
16: \newtheorem{rem}[thm]{Remark}
17: \newtheorem{rems}[thm]{Remarks}
18: \renewcommand{\theequation} {\thesection.\arabic{equation}}
19:
20:
21: \title{$(2+1)$-dimensional topological quantum field theory
22: from subfactors and Dehn surgery formula for 3-manifold invariants}
23:
24: \author{
25: {\sc Yasuyuki Kawahigashi}\footnote{Supported in part by the
26: Grants-in-Aid for Scientific Research, JSPS.}\\
27: Department of Mathematical Sciences\\
28: University of Tokyo, Komaba, Tokyo, 153-8914, JAPAN\\
29: e-mail: {\tt yasuyuki@ms.u-tokyo.ac.jp}\\
30: \vphantom{X}\\
31: {\sc Nobuya Sato}$^*$ \\
32: %\footnote{Supported in part by the
33: %Grants-in-Aid for Scientific Research, JSPS.}\\
34: Department of Mathematics and Information Sciences\\
35: Osaka Prefecture University, Sakai, Osaka, 599-8531, JAPAN\\
36: e-mail: {\tt nobuya@mi.cias.osakafu-u.ac.jp}\\
37: \vphantom{X}\\
38: {\sc Michihisa Wakui}\\
39: Department of Mathematics\\
40: Osaka University, Toyonaka, Osaka, 560-0043, JAPAN \\
41: e-mail: {\tt wakui@math.sci.osaka-u.ac.jp}}
42:
43: \begin{document}
44:
45: \maketitle
46:
47: \begin{abstract}
48: In this paper, we establish the general theory of $(2+1)$-dimensional
49: topological quantum field theory (in short, TQFT) with a Verlinde basis.
50: It is a consequence that we
51: have a Dehn surgery formula for 3-manifold invariants for this kind of
52: TQFT's. We will show that Turaev-Viro-Ocneanu unitary TQFT's obtained from
53: subfactors satisfy the axioms of TQFT's with Verlinde bases. Hence, in a
54: Turaev-Viro-Ocneanu TQFT, we have a Dehn surgery formula for 3-manifolds.
55: It turns out that this Dehn surgery formula is nothing but the formula of
56: the Reshetikhin-Turaev invariant constructed from a tube system,
57: which is a modular category corresponding to the quantum double construction
58: of a $C^*$-tensor category. In the forthcoming paper, we will exbit
59: computations of Turaev-Viro-Ocneanu invariants for several \lq\lq basic
60: 3-manifolds ''. In Appendix, we discuss the relationship
61: between the system of $M_\infty$-$M_\infty$ bimodules arising from the
62: asymptotic inclusion $M \vee M^{op} \subset M_\infty$ constructed from
63: $N \subset M$ and the tube system obtained from a subfactor $N \subset M$.
64: \end{abstract}
65:
66:
67: \section{Introduction}
68: Since the discovery of the celebrated Jones polynomial \cite{J},
69: there have been a great deal of studies on quantum invariants of knots,
70: links, and 3-manifolds. Among them, we deal with the two constructions
71: in this paper; Ocneanu's generalization of the Turaev-Viro
72: invariants \cite{TuraevViro} of 3-manifolds based on triangulation
73: and the Reshetikhin-Turaev invariants \cite{RT} of those based
74: on Dehn surgery. The former uses quantum $6j$-symbols arising from
75: subfactors and the tensor categories we have are not necessarily braided,
76: while the the latter invariants require braiding, or more precisely,
77: modular tensor categories. Methods to construct a modular tensor
78: category from a rational tensor category have been studied by
79: several authors and are often called ``quantum double'' constructions.
80: In the setting of subfactor theory, such a method was first found
81: by Ocneanu using his generalization of the Turaev-Viro topological
82: quantum field theory (TQFT) and Izumi \cite{Izumi1} later gave a formulation
83: based on sector theory.
84:
85: We start with a (rational unitary) tensor category (arising from
86: a subfactor), which is not braided in general.
87: Then we obtain two TQFT's with the methods above as follows. One is
88: a Turaev-Viro-Ocneanu TQFT based on a state sum and triangulation which
89: directly uses $6j$-symbols of the tensor category. The other is
90: a Reshetikhin-Turaev TQFT arising from the modular tensor category
91: we obtain with the quantum double construction of the original category.
92: It is natural to ask what relation we have between these two
93: TQFT's. In order to make a general study on such TQFT's, we first
94: establish a Dehn surgery formula (Proposition \ref{Proposition1})
95: for a general TQFT, which represents an invariant of
96: a 3-manifold in terms of a weight sum of invariants of links that
97: we use for the Dehn surgery construction of the 3-manifold. This
98: formula uses a basis of the Hilbert space for $S^1\times S^1$
99: arising from the TQFT. We show that several nice properties for
100: such a basis are mutually equivalent (Theorem \ref{Theorem6}) and we say
101: that a basis is a {\it Verlinde basis} when such properties hold
102: (Definition \ref{vb}). We next show that
103: a Turaev-Viro-Ocneanu TQFT has a Verlinde basis in this sense
104: (Theorem \ref{verlinde-basis}) and then as a corollary of the Dehn surgery
105: formula, we conclude that the above two TQFT's are identical for
106: any closed 3-manifold (Theorem \ref{main}).
107: (Actually, the claim that a Turaev-Viro-Ocneanu TQFT has a Verlinde
108: basis has been announced by Ocneanu and a proof is presented in
109: \cite[Chapter 12]{EK}, but normalizations are inaccurate
110: there, so we include a proof for this claim for the sake of completeness
111: here, along the line of Ocneanu's tube algebra and braiding arising
112: from the Turaev-Viro-Ocneanu TQFT.)
113: This identity result has been announced by two of us in \cite{SatoWakui}.
114: It has been also announced in page 244 of Ocneanu \cite{O},
115: but it seems to us that his line of
116: arguments are different from ours which relies on our
117: Dehn surgery formula. This result also proves a conjecture
118: in \cite[Section 8.2]{Mueger} and gives an answer to a question
119: in \cite[page 546]{Kerler}.
120: We refer to the book \cite{EK} for subfactor theory and its applications.
121: Throughout this paper, we only consider subfactors with
122: finite depth and finite index.
123:
124: \bigskip
125: \noindent
126: Acknowledgement: A part of this work was done while the first two
127: authors visited Mathematical Sciences Research Institute at Berkeley.
128: We appreciate their financial supports and hospitality.
129:
130:
131: \section{$(2+1)$-dimensional TQFT with Verlinde basis and Dehn surgery
132: formula}
133:
134: Giving a $(2+1)$-dimensional topological quantum field theory $Z$, we have
135: a 3-manifold invariant $Z(M)$ in the canonical way. The purpose of this
136: section is to describe $Z(M)$ in terms of the $S$-matrix and framed links,
137: and to give a criterion for the Verlinde identity \cite{Verlinde} to hold in
138: our framework.
139: \par
140: Throughout this section, we use the following notations.
141: For an oriented manifold $M$, we denote by $-M$ the same manifold with the opposite orientation. For an orientation preserving diffeomorphism $f:M\longrightarrow N$, we denote by $-f$ the same diffeomorphism viewed as an orientation preserving diffeomorphism $-M\longrightarrow -N$. We regard the empty set $\emptyset $ as an oriented closed surface with the unique orientation. The dual vector space of a finite dimensional vector space $V$ over $\mathbb{C}$ is denoted by $V^{\ast }$, and the dual basis of a basis $\{ v_i\} _i$ is denoted by $\{ v_i^{\ast }\} _i$.
142: The closure of a subspace $X$ in the $3$-sphere $S^3$ is denoted by $\overline{X}$.
143:
144: \par \medskip \noindent
145: {\bf Convention of orientations for manifolds}.\
146: \par
147: Throughout this section, we assume that the $3$-sphere $S^3$ is oriented, and assume that any $3$-dimensional submanifold of $S^3$ is oriented by the orientation induced from $S^3$. We also assume that the solid torus $D^2\times S^1$ equips with the orientation such that the diffeomorphism $h:D^2\times S^1\longrightarrow \mathbb{R}^3\subset S^3$ defined by
148: $$h((x,y),e^{i\theta })=((2+x)\cos \theta , (2+x)\sin \theta ,y)$$
149: is orientation preserving, and assume that the torus $S^1\times S^1$ is oriented by the orientation induced from $D^2\times S^1$ (see Figure \ref{Figure1}).
150:
151: \begin{figure}[hbtp]
152:
153: \begin{center}
154: \includegraphics[width=5cm]{sw1.eps}
155: \caption{the orientation of torus $S^1\times S^1$ \label{Figure1}}
156: \end{center}
157: \end{figure}
158:
159: \par \medskip
160: Let us recall the definition of $(2+1)$-dimensional topological quantum field theory due to Atiyah \cite{Atiyah}, \cite{FSS}.
161: By a $3$-cobordism we mean a triple $(M; \Sigma _1,\Sigma _2)$ consisting of a compact oriented $3$-manifold $M$ and two closed oriented surfaces $\Sigma _1, \Sigma _2$ such that $\partial M=(-\Sigma _1)\cup \Sigma _2$ and $\Sigma _1\cap \Sigma _2=\emptyset $. For an oriented closed surface $\Sigma $, the identity cobordism is given by $Id_{\Sigma }=(\Sigma \times [0,1];\Sigma \times \{ 0\} ,\Sigma \times \{ 1\} )$.
162:
163:
164: \par
165: \bigskip
166: \begin{defn}
167: Let $Z$ be a functor consisting of the following three functions.
168: \begin{enumerate}
169: \item[(1)]
170: To each closed oriented surface $\Sigma $, it assigns a finite dimensional $\mathbb{C}$-vector space $Z(\Sigma )$.
171: \item[(2)]
172: To each cobordism $W$, it assigns a $\mathbb{C}$-linear map $Z_W$.
173: \item[(3)]
174: To each orientation preserving diffeomorphism $f$ between closed oriented surfaces, it assigns a $\mathbb{C}$-linear isomorphism $Z(f)$.
175: \end{enumerate}
176: If $Z$ has the following properties, it is called a {\rm $(2+1)$-dimensional
177: topological quantum field theory} (TQFT in short).
178: \begin{enumerate}
179: \item[(i)] $Z$ is functorial with respect to the composition of orientation preserving diffeomorphisms. More precisely,
180: \begin{enumerate}
181: \item[(a)] $Z(g\circ f)=Z(g)\circ Z(f)$
182: for orientation preserving diffeomorphisms $f$ and $g$.
183: \item[(b)] $Z(id_{\Sigma })=id_{Z(\Sigma )}$ for an oriented closed surface $\Sigma $.
184: \end{enumerate}
185: \item[(ii)] $Z$ is fuctorial with respect to the composition of $3$-cobordisms. More precisely,
186: \begin{enumerate}
187: \item[(a)] If two cobordisms $W_1=(M_1; \Sigma _1,\Sigma _2)$ and $W_2=(M_2; \Sigma _2,\Sigma _3)$ are obtained by cutting a cobordism $W=(M;\Sigma _1,\Sigma _3)$ along $\Sigma _2$ in $M=M_1\cup M_2$, then
188: $$Z_W=Z_{W_2}\circ Z_{W_1}.$$
189: \item[(b)] $Z(i_1)^{-1}\circ Z_{Id_{\Sigma }}\circ Z(i_0)=id_{Z(\Sigma )}$ for the identity cobordism on $\Sigma $, where $i_t:\Sigma \longrightarrow \Sigma \times \{ t\} $ \ $(t=0,1)$ are orientation preserving diffeomorphisms defined by $i_t(x)=(x,t),\ x\in \Sigma $.
190: \end{enumerate}
191: \item[(iii)] Let $W=(M,\Sigma _1,\Sigma _2),\ W'=(M',\Sigma _1',\Sigma
192: _2')$ be two cobordisms. Suppose that there is an orientation preserving
193: diffeomorphism $h:M\longrightarrow M'$ such that $h(\Sigma _i)=\Sigma
194: _i'$ for $i=1,2$. Then, for $f_1:=-h\vert _{\Sigma _1}:\Sigma
195: _1\longrightarrow \Sigma _1' \ and \ f_2:=h\vert _{\Sigma _2}:\Sigma _2\longrightarrow \Sigma _2'$, the following diagram commutes.
196: $$\begin{CD}
197: Z(\Sigma _1)@>Z(f_1)>> Z(\Sigma _1') \\
198: @VZ_{W}VV @VVZ_{W'}V \\
199: Z(\Sigma _2) @>>Z(f_2)> Z(\Sigma _2')
200: \notag
201: \end{CD}$$
202: \item[(iv)] Let $W_1=(M;\Sigma _1,\Sigma _2)$, $W_2=(N;\Sigma _2',\Sigma _3)$ be two cobordisms, and $f:\Sigma _2\longrightarrow \Sigma _2'$ an orientation preserving diffeomorphism. Then, for the cobordism $W:=(N\cup _fM; \Sigma _1, \Sigma _3)$ obtained by gluing of $W_1$ to $W_2$ along $f$,
203: $$Z_W=Z_{W_2}\circ Z(f)\circ Z_{W_1}.$$
204: \item[(v)] There are natural isomorphisms
205: \begin{enumerate}
206: \item[(a)] $Z(\emptyset )\cong \mathbb{C}$.
207: \item[(b)] $Z(\Sigma _1\coprod \Sigma _2)\cong Z(\Sigma _1)\otimes Z(\Sigma _2)$ for oriented closed surfaces $\Sigma _1$ and $\Sigma _2$.
208: \item[(c)] $Z(-\Sigma )\cong Z(\Sigma )^{\ast }$ for an oriented closed surface $\Sigma $.
209: \end{enumerate}
210: \end{enumerate}
211: \end{defn}
212:
213: \par \medskip
214: \begin{rems}
215: 1. For an oriented closed $3$-manifold $M$, we have a cobordism $W=(M;\emptyset ,\emptyset )$. This cobordism $W$ induces a linear map
216: $$\mathbb{C}\cong Z(\emptyset )\xrightarrow {Z_W}Z(\emptyset )\cong \mathbb{C}.$$
217: We denote by $Z(M)$ the image of $1$ under the above map. By the condition (iii) we see that $Z(M)$ is a topological invariant of $M$.
218: \par
219: \noindent
220: 2. Let $\varGamma_{\Sigma }$ denote the mapping class group of the oriented closed surface $\Sigma $. Then, we have a representation of $\varGamma_{\Sigma }$
221: $$\rho :\varGamma_{\Sigma }\longrightarrow GL(Z(\Sigma )),\ [f] \longmapsto Z(f),$$
222: where $[f]$ denotes the isotopy class of $f$.
223: \par
224: \noindent
225: 3. For an oriented closed surface $\Sigma $, we have $\dim Z(\Sigma )=Z(\Sigma \times S^1)$.
226: \par
227: \noindent
228: 4. By a $3$-cobordism with parametrized boundary we mean a triple $(M; j_1,j_2)$ consisting of a compact oriented $3$-manifold $M$, an orientation reversing embedding $j_1:\Sigma _1\longrightarrow \partial M$ and an orientation preserving embedding $j_2:\Sigma _2\longrightarrow \partial M$ such that $\partial M=(-j_1(\Sigma _1))\cup j_2(\Sigma _2)$ and $j_1(\Sigma _1)\cap j_2(\Sigma _2)=\emptyset $. Any cobordism with parametrized boundary ${\cal W}=(M; \Sigma _1\xrightarrow {\ j_1\ } \partial M,\Sigma _2 \xrightarrow {\ j_2\ } \partial M)$ induces a linear map $Z_{\cal W}:Z(\Sigma _1)\longrightarrow Z(\Sigma _2)$ such that the following diagram commutes.
229:
230: \begin{equation*}
231: \begin{CD}
232: Z(\Sigma _1) @>Z_{\cal W}>> Z(\Sigma _2) \\
233: @VZ(j_1)VV @VVZ(j_2)V \\
234: Z(j_1(\Sigma _1)) @>Z_W>> Z(j_2(\Sigma _2))
235: \end{CD}
236: \end{equation*}
237:
238: Here, we set $W=(M;j_1(\Sigma _1),j_2(\Sigma _2))$.
239: \end{rems}
240:
241: \par \bigskip
242: \par \bigskip
243: Let $Z$ be a $(2+1)$-dimensional TQFT. We
244: consider the cobordism $W:=(Y\times S^1; \Sigma _1\sqcup \Sigma _2, \Sigma _3)$, where $Y$ is
245: the compact oriented surface in $\mathbb{R}^3$
246: depicted in Figure \ref{Figure5} and $\Sigma _i=C_i\times S^1$ for $i=1,2,3$.
247: Then, $W$ induces a
248: linear map $Z_W: Z(S^1\times S^1)\otimes Z(S^1\times S^1)\longrightarrow
249: Z(S^1\times S^1)$. It
250: can be easily verified that the map $Z_W$ gives an associative algebra structure
251: on $Z(S^1\times S^1)$. The identity element of this algebra is given by
252: $Z_{W_0}(1)$, where $W_0:=(D^2\times S^1; \emptyset , S^1\times S^1)$. We
253: call this algebra the {\it fusion algebra associated with $Z$}.
254:
255: \begin{figure}[hbtp]
256: \begin{center}
257: \scalebox{1.3}{\includegraphics[height=2.5cm]{sw5.eps} }
258: \caption{the compact oriented surface $Y$ \label{Figure5}}
259: \end{center}
260: \end{figure}
261:
262: Let us introduce a Dehn surgery formula of $Z(M)$.
263: Let $Z$ be a $(2+1)$-dimensional TQFT, and $S:S^1\times S^1\longrightarrow S^1\times S^1$ the orientation preserving diffeomorphism defined by $S(z,w)=(\bar{w},z),\ (z,w)\in S^1\times S^1$, where we regard $S^1$ as the set of complex numbers of absolute value $1$. Given a basis $\{ v_i\} _{i=0}^m$ of $Z(S^1\times S^1)$, we define $S_{ji}\in \mathbb{C}$ by $Z(S)v_i=\sum_{j=0}^mS_{ji}v_j$.
264: \par
265: Let $\{ v_i\} _{i=0}^m$ be a basis of $Z(S^1\times S^1)$. Then, we can define a framed link invariant as follows. Let $L=L_1\cup \cdots \cup L_r$ be a framed link with $r$-components in the 3-sphere $S^3$, and $h_i: D^2\times S^1\longrightarrow N(L_i)$ be the framing of $L_i$ for each $i\in \{ 1, \cdots , r\} $, where $N(L_i)$ denotes the tubular neighborhood of $L_i$.
266: We fix an orientation for $\partial N(L_i)$ such that $j_i:=h_i\vert _{\partial D^2\times S^1}:S^1\times S^1\longrightarrow \partial N(L_i)$ is orientation preserving. Since the orientation for $N(L_i)$ is not compatible with the orientation for the link exterior $X:=\overline{S^3-N(L_1)\cup \cdots \cup N(L_r)}$, we can consider the cobordism with parametrized boundary ${\cal W}_L:=(X; \coprod \limits_{i=1}^r j_i ,\emptyset )$.
267: This cobordism induces a $\mathbb{C}$-linear map $Z_{{\cal W}_L}: \bigotimes \limits_{i=1}^rZ(S^1\times S^1)\longrightarrow \mathbb{C}$.
268: It is easy to see that for each $i_1,\cdots ,i_r=0,1,\cdots ,m$ the complex number $J(L; i_1, \cdots , i_r):=Z_{{\cal W}_L}(v_{i_1}\otimes \cdots \otimes v_{i_r})$
269: is a framed link invariant of $L$.
270:
271: \par\bigskip
272: \begin{prop}
273: \label{Proposition1}
274: Let $Z$ be a (2+1)-dimensional TQFT and $\{ v_i\} _{i=0}^m$ a basis of $Z(S^1\times S^1)$ such that $v_0$ is the identity element in the fusion algebra.
275: Let $M$ be a closed oriented 3-manifold obtained from $S^3$ by Dehn surgery along a framed link $L=L_1\cup \cdots \cup L_r$.
276: Then, the 3-manifold invariant $Z(M)$ is given by the formula
277: $$Z(M)=\sum_{i_1, \cdots , i_r=0}^m S_{i_1,0}\cdots S_{i_r,0}J(L;i_1, \cdots , i_r).$$
278: \end{prop}
279:
280: \medskip \noindent
281: {Proof}.\
282: Let $X$ be the link exterior of $L$ in $S^3$, and $h_i:D^2\times S^1\longrightarrow \partial N(L_i)$ the framing of $L_i$ for each $i\in \{ 1,\cdots ,r\} $.
283: Then, by using attaching maps $f_i: S^1\times S^1\longrightarrow \partial N(L_i)$ satisfying $f_i(S^1\times 1)=h_i(1\times S^1)$, $i=1,\cdots ,r$, we have
284: $$M=X\bigcup \nolimits_{\amalg _{i=1}^rf_i}(\coprod^rD^2\times S^1).$$
285: Therefore,
286: $$Z(M)=Z_{W_2}\circ Z(\coprod _{i=1}^rf_i)\circ Z_{W_1},$$
287: where $W_1:=(\coprod^r(D^2\times S^1); \emptyset , \coprod ^r(S^1\times S^1))$ and $W_2:=(X; -\partial X, \emptyset )$. This implies that
288: $$Z(M)=Z_{\mathcal{W}_L}\circ (\bigotimes _{i=1}^rZ(j_i^{-1}))\circ (\bigotimes _{i=1}^rZ(f_i))\circ (\bigotimes ^rZ_{W_0}),$$
289: where $j_i:=h_i\vert _{\partial D^2\times S^1}:S^1\times S^1\longrightarrow \partial N(L_i)$ for $i=1,\cdots ,r$.
290: \par
291: For each $i$ we can choose $f_i$ satisfying $j_i^{-1}\circ f_i=S$ up to isotopy
292: as for $f_i$ satisfying $f_i(S^1\times 1)=h_i(1\times S^1)$.
293: Then, we have
294: $$Z(M)=Z_{\mathcal{W}_L}\circ (\bigotimes _{i=1}^rZ(S))\circ (\bigotimes ^rZ_{W_0}).$$
295: This implies that
296: $$Z(M)=\sum_{i_1, \cdots , i_r=0}^m S_{i_1,0}\cdots S_{i_r,0}J(L;i_1, \cdots , i_r).$$
297: This completes the proof. \hfill Q.E.D.
298: \par \bigskip
299:
300: \begin{figure}[hbtp]
301: \begin{center}
302: \includegraphics[width=4cm]{sw2.eps}
303: \caption{the Hopf link $H$ \label{Figure2}}
304: \end{center}
305: \end{figure}
306:
307: \par \bigskip
308: \begin{prop}
309: \label{Proposition2}
310: Let $Z$ be a $(2+1)$-dimensional TQFT and $\{ v_i\} _{i=0}^m$ a basis of $Z(S^1\times S^1)$.
311: Let $H$ be the Hopf link depicted as in Figure \ref{Figure2}, and $U$ the orientation preserving diffeomorphism from $S^1\times S^1$ to $-S^1\times S^1$ defined by $U(z,w)=(z,\bar{w})$.
312: Then, the following are equivalent.
313: \begin{enumerate}
314: \item[(i)] $J(H;i,j)=S_{ij}$ for all $i,j=0,1,\cdots ,m$.
315: \item[(ii)] $Z(U)v_i=v_{i}^{\ast }$ for all $i=0,1,\cdots ,m$.
316: \end{enumerate}
317: \end{prop}
318:
319: \noindent
320: {Proof}. \
321: Let $h_i:D^2\times S^1\longrightarrow N(H_i)$ be the framing of $H_i$ for each $i=1,2$. Putting $j_i=h_i\vert _{S^1\times S^1}\ (i=1,2)$ and $X:=\overline{S^3-N(H_1)\cup N(H_2)}$, we have a cobordism with parametrized boundary ${\cal W}:=(X; j_1,-j_2)$.
322: Then
323: $$Z_{\cal W}(v_i)=\sum\limits_{i=0}^mJ(H;i,j)v_j^{\ast }.$$
324: Let
325: $V_1$ be the tubular neighborhood of $N(H_1)$ depicted as in Figure \ref{Figure3}.
326:
327: \begin{figure}[htbp]
328: \begin{center}
329: \setlength{\unitlength}{1cm}
330: \includegraphics[height=4cm]{sw3.eps}
331: \caption{ \label{Figure3}}
332: \end{center}
333: \end{figure}
334:
335: Setting $V_2:=\overline{S^3-V_1}$, and defining orientation preserving diffeomorphisms $j_1':S^1\times S^1\longrightarrow \partial V_1$ and $j_2':S^1\times S^1\longrightarrow \partial V_2$ in parallel to $j_1$ and $j_2$ respectively, we see that
336: \par
337: $\bullet$\ $\partial V_1=-\partial V_2$,
338: \par
339: $\bullet$\ $X=\overline{V_1-N(H_1)}\cup \overline{V_2-N(H_2)}$,
340: \par
341: $\bullet$\ ${\cal W}_1:=(\overline{V_1-N(H_1)}; j_1, j_1')\cong Id_{S^1\times S^1}\quad \text{as cobordisms}$,
342: \par
343: $\bullet$\ ${\cal W}_2:=(\overline{V_2-N(H_2)}; -j_2', -j_2)\cong Id_{-S^1\times S^1}\quad \text{as cobordisms}$.
344:
345:
346: Since
347: $$(-j_2')^{-1}\circ j_1':\
348: \begin{cases}
349: m \longmapsto -l, \cr
350: l \longmapsto -m. \end{cases}$$
351: for the meridian $m$ and the longitude $l$ (see Figure \ref{Figure4}), it follows that
352: $$Z(-j_2'{^{-1}}\circ j_1')=Z(U\circ S).$$
353:
354: \begin{figure}[htbp]
355: \begin{center}
356: \setlength{\unitlength}{1cm}
357: \includegraphics[width=5cm]{sw4.eps}
358: \caption{\label{Figure4}}
359: \end{center}
360: \end{figure}
361: Therefore, we obtain
362: $$Z_{\cal W}=Z_{{\cal W}_2}\circ Z(U\circ S)\circ Z_{{\cal W}_1},$$
363: and whence
364: $$\sum\limits_{j=0}^mJ(H;i,j)v_j^{\ast }=Z_{\cal W}(v_i)=(\theta \circ
365: Z(U\circ S))(v_i),$$
366: where $\theta :Z(-S^1 \times S^1) \longrightarrow Z(S^1 \times S^1)^{\ast }$
367: denotes the natural isomorphism.
368: Let $(U_{ij})_{i,j=0,1,\cdots ,m}$ be the presentation matrix of
369: $\theta \circ Z(U):Z(S^1 \times S^1) \longrightarrow Z(S^1\times S^1)^{\ast }$
370: with respect to $\{ v_i \}_{i=0}^m$ and $\{ v_i^{\ast }\}_{i=0}^m$.
371: Then, we have
372: $$(J(H;i,j))_{i,j=0,1,\cdots ,m}=(U_{ij})(S_{ij}).$$
373: Hence, we have
374: \begin{eqnarray*}
375: &(J(H;i,j))_{i,j=0,1,\cdots ,m}=(S_{ij})_{i,j=0,1,\cdots ,m} \notag \\
376: \Longleftrightarrow \ & (U_{ij})=I,\ \text{where $I$ is the identity matrix.} \notag \\
377: \Longleftrightarrow \ & (\theta \circ Z(U))(v_i)=v_{i}^{\ast } \
378: (i=0,1,\cdots ,m). \qquad \qquad \qquad \notag
379: \end{eqnarray*}
380: This completes the proof. \hfill Q.E.D.
381:
382: Let us recall that the mapping class group $\varGamma_{S^1\times S^1}$ of the torus $S^1\times S^1$ is isomorphic to the group
383: $SL_2(\mathbb{Z})$ of integral $2\times 2$-matrices with determinant $1$. It is well-known that this group is generated by $S=\begin{pmatrix} 0 & 1 \\ -1 & 0
384: \end{pmatrix}$ and $T=\begin{pmatrix} 1 & 0 \\ 1 & 1\end{pmatrix}$ with relations
385: $S^4=I,\ (ST)^3=S^2$. The matrices $S$ and $T$ correspond to the orientation preserving diffeomorphisms $S^1\times S^1$ to $S^1\times S^1$ which are defined by $S(z,w)=(\bar{w},z)$ and $T(z,w)=(zw,w),\ (z,w)\in S^1\times S^1$, respectively, where we regard $S^1$ as the set of complex numbers of absolute value $1$ .
386: To define the Verlinde basis, we need one more orientation preserving diffeomorphism $U: S^1\times S^1\longrightarrow - S^1\times S^1$ defined by $U(z,w)=(z,\bar{w})$ for $(z,w)\in S^1\times S^1$ (see Figure \ref{Figure6}).
387:
388: \par \bigskip
389: \begin{defn}\label{vb}\ \
390: Let $Z$ be a $(2+1)$-dimensional TQFT. A basis $\{ v_i\} _{i=0}^m$ of $Z(S^1\times S^1)$ is said to be a {\it Verlinde basis} if it has the following properties.
391: \begin{enumerate}
392: \item[(i)] $v_0$ is the identity element of the fusion algebra associated
393: with $Z$.
394: \item[(ii)]
395: \begin{enumerate}
396: \item[(a)] $Z(S)$ is presented by a unitary and symmetric matrix with respect to the basis $\{ v_i\} _{i=0}^m$.
397: \item[(b)] $Z(S)^2v_0=v_0$, and $Z(S)^2v_i\in \{ v_j\} _{j=0}^m$ for all $i$.
398: \item[(c)] We define $S_{ji}\in \mathbb{C}$ by $Z(S)v_i=\sum_{i=0}^mS_{ji}v_j$. Then,
399: \begin{enumerate}
400: \item[1.] $S_{i0}\not= 0$ for all $i$.
401: \item[2.] $N_{ij}^k:=\sum_{l=0}^m\frac{S_{il}S_{jl}\overline{S_{lk}}}{S_{0l}}\ (i,j,k=0,1,\cdots ,m)$ coincide with the structure constants of the fusion algebra with respect to $\{ v_i\} _{i=0}^m$.
402: \end{enumerate}
403: \end{enumerate}
404: \item[(iii)] $Z(T)$ is presented by a diagonal matrix with respect to the basis $\{ v_i\} _{i=0}^m$.
405: \item[(iv)] $Z(U)v_i=v_{i}^{\ast }$ for all $i$ under the identification $Z(-S^1\times S^1)\cong Z(S^1\times S^1)^{\ast }$.
406: \end{enumerate}
407: We call a {\it pre-Verlinde basis} a basis $\{ v_i\} _{i=0}^m$ of $Z(S^1\times S^1)$ satisfying the four conditions (i), (ii.a), (ii.b) and $Z(U)v_0=v_0^{\ast }$.
408: \end{defn}
409:
410: \par \bigskip
411: \begin{rems}
412: 1. If $\{ v_i\} _{i=0}^m$ is a pre-Verlinde basis, then $S_{i0}$ is a real number for all $i$.
413: \par
414: \noindent
415: 2. A Verlinde basis is unique up to order of elements, since $w_i=S_{0i}\sum_{j=0}^m\overline{S_{ji}}v_j$\ $(i=0,1,\cdots ,m)$ are all orthogonal primitive idempotents in the fusion algebra satisfying $1=w_0+w_1+\cdots +w_m$. This fact follows from that the $S$-matrix diagonalizes the fusion rules in conformal field theory \cite{Verlinde}.
416: \par
417: \noindent
418: 3. The map $\overline{\mathstrut \ \cdot \ }:\{ 0,1,\cdots ,m\} \longrightarrow \{ 0,1,\cdots ,m\} $ defined by $Z(S)^2v_i=v_{\bar{i}}$ is an involution satisfying $\bar{0}=0$.
419: \par
420: \noindent
421: 4. The last condition (iv) was introduced in \cite{Wakui} and modified in \cite{SuzukiWakui}.
422: \end{rems}
423:
424: \begin{figure}[hbtp]
425: \begin{center}
426: \setlength{\unitlength}{1cm}
427: \includegraphics[height=6cm]{sw6.eps}
428: \caption{the action of $SL_2(\mathbb{Z})$ \label{Figure6}}
429: \end{center}
430: \end{figure}
431:
432: \par \bigskip
433: Let us describe some basic results on a $(2+1)$-dimensional TQFT with a
434: pre-Verlinde basis.
435:
436:
437: \par \bigskip
438: \begin{lem}
439: \label{Lemma3}
440: Let $Z$ be a (2+1)-dimensional TQFT and $\{ v_i\} _{i=0}^m$ a pre-Verlinde basis.
441: Then, for the cobordism $W=(-D^2\times S^1; S^1\times S^1, \emptyset )$, we have
442: $Z_W=v_0^{\ast }: Z(S^1\times S^1)\longrightarrow \mathbb{C}$. In particular, $Z(S^2\times S^1)=1$.
443: \end{lem}
444:
445: \par \medskip \noindent
446: {Proof}.\ \
447: By considering the orientation preserving diffeomorphism $\tilde{U}: D^2\times S^1\longrightarrow -D^2\times S^1$ defined by $\tilde{U}(z, w)=(z,\bar{w})$, we have $(Z(\tilde{U}\vert _{\partial D^2\times S^1}))(Z(D^2\times S^1))=Z(-D^2\times S^1)\in Z(-S^1\times S^1)$. Since $\tilde{U}\vert _{\partial D^2\times S^1}=U$ and $Z(D^2\times S^1)=v_0$, it follows that $v_0^{\ast }=Z(-D^2\times S^1)=Z_W$ as elements in $Z(S^1\times S^1)^{\ast }$.
448: \par
449: Next, we prove that $Z(S^2\times S^1)=1$. The $3$-manifold $S^2\times S^1$ is regarded as $(D_{+}^2\times S^1)\cup (-D_{-}^2\times S^1)$, where $D_{+}^2=\{ (x,y,z)\in S^2 \ \vert \ z\geq 0\} ,\ D_{-}^2=\{ (x,y,z)\in S^2 \ \vert \ z\leq 0\} $. Thus, for the cobordisms $W_1:=(D_{+}^2\times S^1; \emptyset, \partial D_{+}^2\times S^1)$ and $W_2:=(-D_{-}^2\times S^1; \partial D_{-}^2\times S^1,\emptyset )$, we have
450: $$Z(S^2\times S^1)=Z_{W_2}\circ Z_{W_1}.$$
451: Since $W_1\cong (D^2\times S^1;\emptyset ,S^1\times S^1)=W_0$ and $W_2\cong (-D^2\times S^1;S^1\times S^1,\emptyset )=W$, it follows that
452: $$Z(S^2\times S^1)=Z_{W}\circ Z_{W_0}=v_0^{\ast }(v_0)=1.$$
453: This completes the proof. \hfill Q.E.D.
454:
455: \par \bigskip
456: \begin{lem}
457: \label{Lemma4}
458: Let $Z$ be a (2+1)-dimensional TQFT and $\{ v_i\} _{i=0}^m$ a pre-Verlinde basis. Then,
459: $J(\text{\LARGE $\bigcirc $};i)=S_{i0}$ for each $i\in \{ 0,1,\cdots ,m\} $.
460: \end{lem}
461:
462: \par \medskip
463: \noindent
464: {Proof}. \
465: Let $K$ be the trivial knot with $0$-framing given by unit circle in $\mathbb{R}^3$, and $X=\overline{S^3-N(K)}$ the knot exterior of $K$.
466: The cobordism $W_K=(X; \partial N(K), \emptyset )$ is isomorphic to the cobordism $W:=(-D^2\times S^1; S^1\times S^1, \emptyset )$ via the orientation preserving diffeomorphism $f:X\longrightarrow -D^2\times S^1$ such that
467: $f(M)=l$ and $f(L)=-m$, where $M$ and $L$ are simple closed curves on $\partial N(K)$ depicted in Figure \ref{Figure7}.
468:
469:
470: \begin{figure}[htbp]
471: \vspace{0.5cm}
472: \begin{center}
473: \setlength{\unitlength}{1cm}
474: \scalebox{0.9}[0.9]{\includegraphics[height=6cm]{sw7.eps}}
475: \caption{\label{Figure7}}
476: \end{center}
477: \end{figure}
478:
479: \par
480: Then, we have
481: $$Z_{W_K}=Z_{W}\circ Z(f\vert _{\partial N(K)}).$$
482: Since
483: $$Z(f\vert _{\partial N(K)})\circ Z(j)=Z(S)$$
484: for $j:=h\vert _{S^1\times S^1}$, where $h:D^2\times S^1\longrightarrow N(K)$ is the framing of $K$,
485: we have
486: $$J(K,i)=\langle v_0^{\ast },\ Z(S)(v_i)\rangle =S_{i0}.$$
487: This completes the proof. \hfill Q.E.D.
488:
489: \par \bigskip
490: The framed link invariants $J(L;i_1,\cdots ,i_r)$, $(i_1, \dots, i_r=0,1,\dots,
491: m)$ have the following nice properties.
492:
493: \par
494: \begin{figure}[hbtp]
495: \begin{center}
496: \setlength{\unitlength}{1cm}
497: \includegraphics[height=2cm]{sw8.eps}
498: \caption{a positive curl and a negative curl \label{Figure8}}
499: \end{center}
500: \end{figure}
501:
502: \par \bigskip
503: \begin{lem}
504: \label{Lemma5}
505: Let $Z$ be a $(2+1)$-dimensional TQFT and $\{ v_i\} _{i=0}^m$ a pre-Verlinde basis.
506: For a framed link $L=L_1\cup L_2\cup \cdots \cup L_r$, the invariant $J(L; i_1, \cdots , i_r)$ has the following properties.
507: \par
508: (1) Let $L'=L_1\cup \cdots \cup (-L_k)\cup \cdots \cup L_r$ be the framed link obtained from $L$ by changing the orientation for the $k$-th component $L_k$. Then,
509: $$J(L;i_1,\cdots ,i_k,\cdots ,i_r)=J(L';i_1, \cdots ,\bar{i}_k,\cdots ,i_r).$$
510: \indent
511: (2) Suppose that $\{ v_i\} _{i=0}^m$ satisfies the condition (iii) in the definition of Verlinde basis. We define $t_i\in \mathbb{C}\ (i=0,1,\cdots ,r)$ by $Z(T)v_i=t_iv_i$.
512: Let $L'$ be the framed link obtained from $L$ such that it is same as $L$ except for a small segment of the $k$-th component $L_k$ and the small segment is replaced by a positive or negative curl as shown in Figure \ref{Figure8}. Then, the following equations hold.
513: \par
514: If the small segment is replaced by a positive curl, then
515: $$J(L'; i_1,\cdots ,i_r)=t_{i_k}^{-1}J(L;i_1,\cdots ,i_r).$$
516: If the small segment is replaced by a negative curl, then
517: $$J(L'; i_1,\cdots ,i_r)=t_{i_k}J(L;i_1,\cdots ,i_r).$$
518: \end{lem}
519:
520: \noindent
521: {Proof}.\
522: Without loss of generality, we may suppose that $k=1$.
523:
524: \noindent
525: (1) Let $h_i:D^2\times S^1\longrightarrow N(L_i)$ be the framing of $L_i$ for each $i=1,\cdots ,r$, and $h_1':D^2\times S^1\longrightarrow N(-L_1)$ the framing of $-L_1$. We set $j_i:=h_i\vert _{S^1\times S^1}$ for each $i=1,\cdots ,r$, and $j_1':=h_1'\vert _{S^1\times S^1}$.
526: Then,
527: $$j_1'=j_1\circ S^2,$$
528: since $h_1'=h_1\circ f$ for the orientation preserving diffeomorphism $f:D^2\times S^1\longrightarrow D^2\times S^1$ defined by $f(z,w)=(\bar{z},\bar{w})$ (see Figure \ref{Figure9}).
529:
530: \begin{figure}[hbtp]
531: \begin{center}
532: \setlength{\unitlength}{1cm}
533: \includegraphics[height=5cm]{sw9.eps}
534: \caption{\label{Figure9}}
535: \end{center}
536: \end{figure}
537:
538: Let $X$ be the link exterior of $L$.
539: Since the two cobordisms
540: $${\cal W}_L=(X; \coprod_{i=1}^rj_i, \emptyset ),\quad
541: {\cal W}_{L'}=(X;j_1'\coprod (\coprod _{i=2}^rj_i), \emptyset )$$
542: are isomorphic via the identity map $id_X$, we see that
543: $$Z_{{\cal W}_{L'}}\circ (Z(S^2)\circ id \otimes \cdots \otimes id)=Z_{{\cal W}_L}.$$
544: Since $Z(S^2)v_i=v_{\bar{i}}$, we have
545: $$J(L;i_1,i_2,\cdots ,i_r)=J(L';\bar{i}_1, i_2,\cdots ,i_r).$$
546: \par
547: (2) We suppose that $L'$ arises from $L$ by replacing a small segment of the first component $L_1$ by a positive curl.
548: Let $h_1':D^2\times S^1\longrightarrow N(L_1')$ be the framing of $L_1'$. We set $j_1':=h_1'\vert _{S^1\times S^1}$.
549:
550: \begin{figure}[hbtp]
551: \begin{center}
552: \setlength{\unitlength}{1cm}
553: \includegraphics[height=6cm]{sw10.eps}
554: \caption{\label{Figure10}}
555: \end{center}
556: \end{figure}
557:
558: Let $X'$ be the link exterior of $L'$.
559: Then, there exists an orientation preserving diffeomorphism
560: $f:X'\longrightarrow X$
561: such that
562: \begin{enumerate}
563: \item[(i)] $j_1^{-1}\circ (-f\vert _{\partial N(L_1')})\circ j_1'=T^{-1}$ up
564: to isotopy,
565: \item [(ii)] $-f\vert _{\partial N(L_i)}=id_{\partial N(L_i)}$ for all
566: $i=2,\cdots ,r$.
567: \end{enumerate}
568:
569: This map $f$ gives rise to the isomorphism between
570: the cobordisms ${\cal W}_{L'}=(X';\emptyset ,j_1'\coprod (\coprod
571: _{i=2}^rj_i))$ and ${\cal W}_L=(X;\emptyset ,\coprod _{i=1}^rj_i)$. Hence,
572: we have
573: $$Z_{{\cal W}_{L'}}\circ (Z({j_1'}^{-1})\otimes (\bigotimes
574: _{i=2}^rZ(j_i^{-1})))=Z_{{\cal W}_L} \circ (\bigotimes
575: _{i=1}^rZ(j_i^{-1})\circ Z(-f\vert _{\partial N(L_i)})).$$
576: It follows that
577: $$Z_{{\cal W}_{L'}}=Z_{{\cal W}_L} \circ (Z(T^{-1})\otimes id \cdots \otimes
578: id ).$$
579: This implies that
580: $$J(L'; i_1,\cdots ,i_r)=t_{i_1}^{-1}J(L;i_1,\cdots ,i_r).$$
581:
582: Thus, the proof of the first equation of (2) is completed.
583: By a similar argument, the second equation of (2) can be proved.
584: This completes the proof. \hfill Q.E.D.
585:
586: \begin{figure}[htpb]
587: \begin{center}
588: \setlength{\unitlength}{1cm}
589: \includegraphics[height=5cm]{sw11.eps}
590: \caption{\label{Figure11}}
591: \end{center}
592: \end{figure}
593:
594: \par \bigskip
595: Let us introduce a criterion for the Verlinde identity (ii.c.2).
596: \par
597: We suppose that a $(2+1)$-dimensional TQFT $Z$ arises from a semisimple ribbon $\mathbb{C}$-linear Ab-category in the sense of Turaev \cite{Turaev}. Let $\{ v_i\} _{i=0}^m$ be a pre-Verlinde basis satisfying the equivalent conditions (1) and (2) in Proposition \ref{Proposition2}. We suppose $S_{i0}\not= 0$ for all $i=0,1,\cdots ,m$.
598: \par
599: For the $3$-component framed link $L$ presented by the diagram as in Figure \ref{Figure11},
600: the framed link invariant $J(L;i,j,k)$ coincides with the quantum trace of $J(T_{i,l})\circ J(T_{j,l})$, where $T_{i,l}$ is the colored framed tangle presented by the diagram as in Figure \ref{Figure12}.
601: \par
602: Since $J(T_{i,l})$ is a map from the simple object $l$ to $l$, the invariant $J(T_{i,l})$ is a scalar multiple by $id_l$. We define $a_{i}$ to be this
603: scalar. Since the quantum trace of $T_{i,l}$ is the Hopf link with colors $i,l$, we see that
604: $$a_i\cdot J(\text{\LARGE $\bigcirc $};l)=S_{il}.$$
605: This implies that
606:
607: \begin{equation}
608: J(L;i,j,l)=\dfrac{S_{il}S_{jl}}{S_{l0}}. \label{eq1}
609: \end{equation}
610:
611: \begin{figure}[hbtp]
612: \begin{center}
613: \setlength{\unitlength}{1cm}
614: \scalebox{0.9}[0.9]{\includegraphics[width=3.5cm]{sw12.eps}}
615: \caption{the colored $1-1$ tangle $T_{i,l}$ \label{Figure12}}
616: \end{center}
617: \end{figure}
618:
619: \par
620: On the other hand, by fusing $L_i$ and $L_j$ we have
621:
622: \begin{equation}
623: J(L;i,j,l)=\sum\limits_{i=0}^mN_{ij}^kS_{kl},\label{eq2}
624: \end{equation}
625: where $N_{ij}^k\ (i,j,k=0,1,\cdots ,m)$ are the structure constants of the fusion algebra with respect to $\{ v_i\}_{i=0}^m$.
626: From (\ref{eq1}) and (\ref{eq2}), it follows that
627: $$\sum\limits_{i=0}^mN_{ij}^kS_{kl}=\dfrac{S_{il}S_{jl}}{S_{l0}}.$$
628: Since the above equation induces the Verlinde identity
629: $$N_{ij}^k=\sum_{l=0}^m\frac{S_{il}S_{jl}\overline{S_{lk}}}{S_{l0}},$$
630: the condition (2) in Proposition \ref{Proposition2} implies the condition (ii.c.2) in the definition of Verlinde basis (see \cite{Takata}, \cite{Witten} for similar arguments).
631: If $\{ v_i\} _{i=0}^m$ satisfies the condition (iii) in the definition of Verlinde basis, then we see that the converse is true by using the technique in
632: \cite{Kohno} (see Figure \ref{Figure13}).
633:
634: \begin{figure}[htbp]
635: \begin{center}
636: \setlength{\unitlength}{1cm}
637: \scalebox{0.9}[0.9]{\includegraphics[height=3cm]{sw13.eps}}
638: \caption{\label{Figure13}}
639: \end{center}
640: \end{figure}
641:
642: Thus, we conclude the following.
643:
644: \par \bigskip
645: \begin{thm}
646: \label{Theorem6}
647: Suppose that a $(2+1)$-dimensional TQFT $Z$ arises from a semisimple ribbon $\mathbb{C}$-linear Ab-category.
648: Let $\{ v_i\} _{i=0}^m$ be a pre-Verlinde basis satisfying $S_{i0}\not= 0$ for all $i=0,1,\cdots ,m$.
649: If $Z(U)v_i=v_{i}^{\ast }$ for all $i=0,1,\cdots ,m$, then
650: $N_{ij}^k:=\sum_{l=0}^m\frac{S_{il}S_{jl}\overline{S_{lk}}}{S_{0l}}\ (i,j,k=0,1,\cdots ,m)$ coincide with the structure constants of the fusion algebra with respect to $\{ v_i\} _{i=0}^m$.
651: \par
652: Furthermore, if $\{ v_i\} _{i=0}^m$ satisfies the condition (iii) in the definition of Verlinde basis, then the converse is true. Therefore, the following are equivalent.
653: \begin{enumerate}
654: \item[(i)] $J(H;i,j)=S_{ij}$ for all $i,j=0,1,\cdots ,m$.
655: \item[(ii)] $Z(U)v_i=v_{i}^{\ast }$ for all $i=0,1,\cdots ,m$.
656: \item[(iii)] $N_{ij}^k:=\sum_{l=0}^m\frac{S_{il}S_{jl}\overline{S_{lk}}}{S_{0l}}\ (i,j,k=0,1,\cdots ,m)$ coincide with the structure constants of the fusion algebra with respect to $\{ v_i\} _{i=0}^m$.
657: \end{enumerate}
658: Here, $H$ is the Hopf link presented by the diagram as in Figure \ref{Figure2}.
659: \end{thm}
660:
661:
662: \section{Turaev-Viro-Ocneanu $(2+1)$-dimensional TQFT (Review)}
663:
664: \subsection{Sectors and finite system $\Delta$}
665:
666: For a detailed exposition about sectors, see \cite{Iz}.
667: Let $N \subset M$ be an inclusion of infinite factors. In this case,
668: we also have a similar concept to Jones index named Kosaki
669: index, and if it takes the minimal value, we write it $[M:N]_0$ as in
670: the case of Jones index. In the sequel, we assmue subfactors have
671: finite minimal indices. For the inclusion of $\rho(M)
672: \subset M$, where $\rho \in {\rm End}(M)$, we call $[M:\rho(M)]_0^{1/2}$
673: the statistical dimension of $\rho$ and denote it by $d(\rho)$.
674: We denote the set of $*$-homomorphisms from $N$ to $M$ with finite statistical
675: dimensions by ${\rm Mor}(N, M)_0$.
676: We say that $\rho_1$, $\rho_2 \in {\rm Mor}(N, M)_0$
677: are equivalent if there exisits a unitary $u$ in $M$ such that $u\rho_1(x) =
678: \rho_2(x) u$ for all $x \in N$. This gives an equivalence relation in
679: ${\rm Mor}(N, M)_0$. The set of the equivalencce classes is denoted by
680: ${\rm Sect}(N,M)$ and its element is called an $M$-$N$ {\it sector}, which
681: is denoted by $[\rho]$ for $\rho \in {\rm Mor}(N, M)_0$. In $M$-$M$ sectors
682: ${\rm Sect}(M,M)=:{\rm Sect}(M)$, we have the product
683: $[\rho_1] \cdot [\rho_2] = [\rho_1 \circ \rho_2]$ and the summation $[\rho_1]
684: \oplus [\rho_2]$. (See \cite{Iz} for the definition of the summation of
685: sectors.) These operations define a semiring structure in
686: ${\rm Sect}(M)$. For $[\rho_1], [\rho_2] \in {\rm Sect}(M)$, we define
687: \[
688: (\rho_1, \rho_2)=\{ V \in M| V\rho_1(x) = \rho_2(x) V, \forall x \in M \}.
689: \]
690: It is called the {\it intertwiner space} between $\rho_1$ and $\rho_2$.
691: If $(\rho, \rho) = \mathbb{C}1_M$, we say that $\rho$ is {\it irreducible}.
692: The intertwiner space $(\rho_1, \rho_2)$ has an inner product $\langle V, W
693: \rangle=W^* \cdot V$, $V, W \in (\rho_1, \rho_2)$, if $\rho_1$ is
694: irreducible.
695:
696: Moreover, ${\rm Sect}(M)$ has a conjugation $\overline{[\rho]}=
697: [\bar{\rho}]$. Namely, for $[\rho] \in {\rm Sect}(M)$, there exist an
698: endomorphism $\bar{\rho} \in {\rm End}(M)$ and a pair of intertwiners
699: $R_\rho \in (id, \bar{\rho} \rho)$ and $\bar{R}_\rho \in
700: (id, \rho \bar{\rho})$ such that $\bar{R}_\rho^* \rho(R_\rho)=
701: R_\rho^* \bar{\rho}(\bar{R}_\rho)=1/d(\rho)$. With these operations,
702: ${\rm Sect}(M)$ becomes a $*$-semiring over $\mathbb{C}$.
703:
704: An important thing is that ${\rm Sect}(M)$ is closed under the operations such
705: as product, direct sum, irreducible decommposition and conjugation.
706:
707: Let us introduce the notion of a finite system $\Delta$ of ${\rm End}(M)_0$,
708: which is a basic data to describe a topological quantum field theory from
709: subfactors. Since the embedding $\iota : N \hookrightarrow M$ is an element of
710: ${\rm Mor}(N, M)_0$, we can consider the sector $[\iota] \in {\rm Sect}(N, M)$.
711: We note that the conjugation $\overline{[\iota]}$ is an element of
712: ${\rm Sect}(M, N)$, and the product $[\iota] \overline{[\iota]}$ becomes an
713: element of ${\rm Sect}(M)$.
714: In a similar way, $\overline{[\iota]} [\iota]$ becomes an element of
715: ${\rm Sect}(N)$. By decomposing $([\iota] \overline{[\iota]})^n$, $([\iota]
716: \overline{[\iota]})^n [\iota] $, $\overline{[\iota]} ([\iota]
717: \overline{[\iota]})^n $ and $(\overline{[\iota]} [\iota])^n$ into irreducible
718: sectors, we get $M$-$M$, $M$-$N$, $N$-$M$ and $N$-$N$ sectors
719: responsibly. (A sector $[\rho]$ is said to be {\it irreducible} if
720: $\rho$ is irreducible.) If the number of the irreducible sectors in the above
721: decompositions is finite, then the subfactor $N \subset M$ is called of
722: {\it finite depth}. Throughout this paper, we only consider subfactors with
723: finite depth and finite index. For a finite depth subfactor, we
724: get finitely many irrducible $M$-$M$ sectors. In other words, we have a
725: representative set $\Delta=\{ \rho_\xi \}_{\xi \in \Delta_0}$ of finitely many
726: irreducible $M$-$M$ sectors such that \\
727: (i) $[\rho_\xi]=[\rho_\eta]$ if and only if $\xi=\eta$ \\
728: (ii) There exists $e \in \Delta_0$ such that $\rho_e =id$ \\
729: (iii) For any $\xi \in \Delta_0$, there exists $\bar{\xi} \in \Delta_0$
730: such that $\overline{[\rho_\xi]}=[\rho_{\bar{\xi}}]$ \\
731: (iv) There exist non-negative intetgers $N_{\xi,\eta}^\zeta$ such that
732: $[\rho_\xi][\rho_\eta]=\oplus_{\zeta \in \Delta_0} N_{\xi,\eta}^\zeta
733: [\rho_\zeta]$. \\
734:
735: \noindent
736: We call this $\Delta$ a finite system of ${\rm End}(M)_0$ or simply a
737: {\it finite system}.
738:
739: \begin{rem}
740: We note that $\Delta$ can be seen as a $C^*$-tensor category in the following
741: manner. The objects of the category are $\mathbb{C}$-linear span of elements of
742: $\Delta$, the morphisms of the category are intertwiners, and the
743: tensor product structure is given by the compositions of endomorphisms
744: in objects of the category. By an abuse of notation, we denote this
745: category by $\Delta$. See \cite{LR} for the details.
746: \end{rem}
747: \subsection{Turaev-Viro-Ocneanu TQFT}
748:
749: We need some preparations before constructing the Turaev-Viro-Ocneanu
750: TQFT. In the sequel, we write $\xi$ instead of $\rho_\xi$ and so forth,
751: for simplicity.
752:
753: Let $\Delta$ be the finite system of ${\rm End}(M)_0$. We consider the following
754: diagram.
755: \[
756: \begin{CD}
757: \xi \cdot \alpha \cdot \eta
758: @< \xi(T_1) << \xi \cdot \beta \\
759: @A{T_3}AA @AA{T_2}A \\
760: \gamma \cdot \eta
761: @< T_4 << \delta
762: \end{CD}
763: \]
764: Here, $\alpha, \beta, \gamma, \xi, \eta, \delta
765: \in \Delta$ and $T_1 \in (\beta, \alpha \cdot \eta)$,
766: $T_2 \in (\delta, \xi \cdot \beta)$, $T_3 \in (\gamma,
767: \xi \cdot \alpha)$, $T_4 \in (\delta, \gamma \cdot \eta)$.
768: Then, the composition of intertwiners $T_4^* \cdot T_3^* \cdot \xi(T_1)
769: \cdot T_2$ belongs to $(\delta, \delta)$. Since $\delta$ is assumed
770: to be irreducible, this composition of intertwiners is regarded as
771: a complex number. We call this number a {\it quantum $6j$-symbol}.
772:
773: The above diagram can be seen as a tetrahedron as in Figure \ref{6j}.
774: \begin{figure}[htbp]
775: \begin{center}
776: \scalebox{0.5}{
777: \includegraphics{6j.eps}}
778: \end{center}
779: \caption{A tetrahedron as a diagram}
780: \label{6j}
781: \end{figure}
782:
783: We assign $d(\beta)^{-1/2}d(\gamma)^{-1/2} T_4^* \cdot T_3^* \cdot \xi(T_1)
784: \cdot T_2$ to this tetrahedron.
785:
786: Let $V$ be an oriented closed 3-dimensional manifold. Choose a triangulation
787: of $V$ and write it ${\cal T}$. To each vertex in ${\cal T}$, we assign
788: the factor $M$, to each edge in ${\cal T}$ an element in $\Delta$ and to each face in
789: ${\cal T}$ an intertwiner.
790: Let $E$ be the set of the edges in ${\cal T}$, $e$ be an assignment of
791: elements in $\Delta$ to the edges in ${\cal T}$, and $\varphi$ be an
792: assignment of intertwiners to the faces in ${\cal T}$.
793: A tetrahedron $\tau$ has the labeled edges by the assignment $e$ and
794: the labeled faces by the assignment $\varphi$. Hence, to a tetrahedron
795: $\tau$, we assign a complex number defined by using the $6j$-symbol as above
796: or its complex conjugate depending on the orientation of $\tau$.
797: We denote this complex number by $W(\tau;e,\varphi)$. We multiply these
798: $W$'s and the weights $\prod_E d(\xi)$. Then, sum up all these resulting
799: values. Finally, we multiply some weights coming from the vertices
800: in ${\cal T}$.
801: \[
802: Z_\Delta(V,{\cal T})= \lambda^{-a} \sum_e(\prod_E d(\xi))
803: \sum_\varphi \prod_\tau W(\tau;e,\varphi),
804: \]
805: where $\lambda=\sum_{\xi \in \Delta_0} d(\xi)^2$ and $a$ is
806: the number of the vertices in ${\cal T}$. This $Z_\Delta(V,{\cal T})$ is
807: proven to be independent of any choice of triangulations because of
808: the properties
809: of quantum $6j$-symbols. (See \cite{EK} for a detailed account.) Namely,
810: $Z_\Delta(V,{\cal T})$ turns out to be a topological invariant of $V$. So, we
811: drop ${\cal T}$ off from $Z_\Delta(V,{\cal T})$ and denote this value
812: by $Z_\Delta(V)$. We call $Z_\Delta(V)$ the {\it Turaev-Viro-Ocneanu invariant}
813: of the 3-dimensional manifold $V$.
814: When $V$ is an oriented, compact 3-dimensional manifold possibly with
815: boundary, first we fix a triangulation $\partial {\cal T}$ of the
816: boundary of $V$ and extend it to the whole triangulation of $V$. Then,
817: we assign an element in $\Delta$ to each edge in $\partial {\cal T}$ and
818: assign an intrertwiner to each face. We fix these assignments to the end
819: and denote them by $\partial e$ and $\partial \varphi$, respectively. In a similar
820: way to the closed case, we assign $M$, an element in $\Delta$ and an
821: intertwiner to each vertex, each edge and each face in the triangulation
822: ${\cal T} \setminus \partial {\cal T}$. We make
823: $Z_\Delta(V,{\cal T}, \partial e, \partial \varphi)$ as above:
824: \[
825: Z_\Delta(V,{\cal T}, \partial e, \partial \varphi)=
826: \lambda^{-a+\partial a/2} \prod_{\partial e}
827: d(\xi)^{1/2} \sum_{e \setminus \partial e}(\prod_{E \setminus \partial E}
828: d(\xi)) \sum_\varphi \prod_\tau W(\tau;e,\varphi),
829: \]
830: where $\partial a$ is the number of the vertices on the boundary. This
831: value $Z_\Delta(V,{\cal T}, \partial e, \partial \varphi)$ does not
832: depend on the assignments of the vertices, the edges and the faces in
833: ${\cal T} \setminus \partial {\cal T}$. Such extended
834: $Z_\Delta$ gives rise to a unitary TQFT because to an oriented closed surface,
835: it assigns a finite dimensional Hilbert space with the inner product
836: induced from the space of intertwiners. (See \cite{EK} for the detailed
837: construction.) We call this TQFT the {\it $(2+1)$-dimensional
838: Turaev-Viro-Ocneanu TQFT} and denote it by $Z_\Delta$, again.
839:
840:
841: \section{Verlinde basis of $Z_{\Delta}(S^1 \times S^1)$}
842:
843: Let $N \subset M$ be a subfactor of an infinite factor $M$ with finite
844: index and finite depth, and let $\Delta$ be a finite system of irreducible
845: $M$-$M$ endomorphisms arising from the subfactor. We write, for instance,
846: $\xi$ instead of $\rho_\xi$ and so forth for the elements of $\Delta$.
847:
848: Based on $\Delta$, we construct a new finite dimensional $C^*$-algebra named
849: the {\it tube algebra} ${\rm Tube} \Delta$ as in \cite{Ocneanu}. (Also see
850: \cite{Izumi1}, but our normalization convention is different from that
851: there.) In this section, ${\rm Tube} \Delta$ plays a crutial role to find a
852: nicely behaved basis of $Z_{\Delta}(S^1 \times S^1)$, which we call a Verlinde
853: basis. It makes the Turaev-Viro-Ocneanu TQFT a rich theory.
854:
855: We sometimes use the simple notation $Z(S^1 \times S^1)$ instead of
856: $Z_{\Delta}(S^1 \times S^1)$ in the sequel.
857:
858: \subsection{Tube algebras}
859:
860: A tube algebra ${\rm Tube}\Delta$, which was first introduced by Ocneanu
861: in \cite{Ocneanu}, is defined by $\bigoplus_{\xi,\eta,\zeta} (\xi \cdot
862: \zeta, \zeta \cdot \eta)$ as a vector space over $\mathbb{C}$. Its element
863: is a linear combination of the composition $T_2 \cdot T_1^*$ of the orthonormal
864: bases of intertwiner spaces $T_1 \in (\delta, \xi \cdot \zeta)$, $T_2 \in (\delta,
865: \zeta \cdot \eta)$, where $\xi, \eta, \zeta$ and $\delta$ run over $\Delta_0$.
866: An element in ${\rm Tube}\Delta$ can be depicted as in the left-hand side of
867: Figure \ref{tube}. We will define a product structure and a $*$-structure on it.
868:
869:
870: \noindent
871: \underline{The product structure} \\
872: Let $X=X_2 X_1^* \in (\xi \cdot \zeta, \zeta \cdot \eta)$ for $X_1 \in
873: (\delta, \xi \cdot \zeta)$, $X_2 \in (\delta, \zeta \cdot \eta)$ and
874: $Y=Y_2 Y_1^* \in (\xi' \cdot \zeta', \zeta' \cdot \eta')$ for $Y_1 \in
875: (\delta', \xi' \cdot \zeta')$, $Y_2 \in (\delta', \zeta' \cdot \eta')$.
876: Then, the product of $X$ and $Y$ in ${\rm Tube} \Delta$ is defined by
877: the following formula.
878: \[
879: X \cdot Y = \delta_{\eta, \xi'} \lambda \sum
880: d(\xi)^{-1/2} d(\eta')^{-1/2} d(\eta) \lambda(X,Y;Z)Z.
881: \]
882: Here, the summation is taken over $Z=Z_2 Z_1^*$, $Z_1 \in (\tau, \xi \cdot \nu)$,
883: $Z_2 \in (\tau, \nu \cdot \eta')$, where $Z_1$ and $Z_2$ are orthonormal bases of
884: the intertwiner spaces $(\tau, \xi \cdot \nu)$ and $(\tau, \nu \cdot
885: \eta')$ respectively, and $\lambda(X,Y;Z)$ is the value of $Z_\Delta$ of
886: the 3-manifold depicted in Figure \ref{tube-c}.
887: For general elements $X$, $Y$ in ${\rm Tube} \; \Delta$, we
888: define the product of them by linearlity since $X$ and $Y$ are linear
889: combinations of the forms $X_2 X_1^*$ and $Y_2 Y_1^*$ as above.
890:
891: \begin{figure}[htbp]
892: \begin{center}
893: \scalebox{0.35}{
894: \includegraphics{tube-c.eps}}
895: \end{center}
896: \caption{A coeffcient of the product of $X \cdot Y$}
897: \label{tube-c}
898: \end{figure}
899:
900: \noindent
901: \underline{The $*$-structure} \\
902: Let $X$ be as above. Then, we can consider $X$ as a tube as in the right-hand
903: side of Figure \ref{tube}.
904: The $*$-operation is defined by the inversing the tubes inside out. We denote
905: this $*$-operation by $X^*$. See Figure \ref{tube-r}.
906:
907: \begin{figure}[htbp]
908: \begin{center}
909: \scalebox{0.5}{
910: \includegraphics{tube.eps}}
911: \end{center}
912: \caption{An element of the tube algebra as a tube}
913: \label{tube}
914: \end{figure}
915:
916: \begin{figure}[htbp]
917: \begin{center}
918: \scalebox{0.4}{
919: \includegraphics{tube-r.eps}}
920: \end{center}
921: \caption{An inversed tube}
922: \label{tube-r}
923: \end{figure}
924:
925: With the product and the $*$-structure defined as above, ${\rm Tube} \Delta$
926: becomes a finite dimensional $C^*$-algebra. Since any finite dimensional
927: $C^*$-algebra is semisimple, we may assume that
928: ${\rm Tube} \Delta \cong \bigoplus_{i=0}^r M_{n_i}({\mathbb C})$.
929:
930: We observe that the definition of tube algebras is compatible
931: with the operations of Turaev-Viro-Ocneanu TQFT such as gluing,
932: cutting and so forth.
933:
934: \begin{rem}
935: In \cite{Izumi1}, M. Izumi has introduced the tube algebra in the
936: setting of sectors, but it is slightly different from ours in the
937: normalization coefficients. We followed the definition of the tube
938: algebra in \cite[Chapter 12]{EK}.
939: \end{rem}
940:
941: Before we start the analysis of tube algebras, we list the notations that
942: we will use frequently.
943:
944: For the solid torus $D^2 \times S^1$, we denote
945: the value of $Z_\Delta$ of $D^2 \times S^1$ assigned a vector $\lambda$
946: on the boundary by $Z_\Delta(D^2 \times S^1;\lambda)$.
947:
948: For the 3-manifold $D^2 \times S^1 \setminus {\rm Int}D_0^2 \times S^1$,
949: where $D_0$ is contained in ${\rm Int} D^2$, we denote the value of $Z_\Delta$ of
950: $D^2 \times S^1 \setminus {\rm Int}D_0^2 \times S^1$ assigned vectors $\lambda$ and
951: $\mu$, $\lambda$ for the boundary of $D_0^2 \times S^1$ and $\mu$ for the boundary of
952: $D^2 \times S^1$ by $Z_\Delta(D^2 \times S^1 \setminus {\rm Int}D_0^2 \times S^1;
953: \lambda,\mu)$. See Figure \ref{solid-a}.
954:
955: \begin{figure}[htbp]
956: \begin{center}
957: \scalebox{0.4}{
958: \includegraphics{solid1.eps}}
959: \end{center}
960: \caption{The labeled 3-manifold obtained by removing another solid torus from a solid torus}
961: \label{solid-a}
962: \end{figure}
963:
964: In a similar manner, for the 3-manifold $D^2 \times S^1 \setminus ( {\rm Int} D_1^2
965: \times S^1 \cup {\rm Int} D_2^2 \times S^1)$, where $D_1^2$ and $D_2^2$ are two disjoint
966: disks contained in ${\rm Int} D^2$, we denote the value of $Z_\Delta$ of $D^2 \times S^1
967: \setminus ({\rm Int} D_1^2 \times S^1 \cup {\rm Int} D_2^2 \times S^1)$
968: assigned vectors $\lambda$, $\mu$ and $\nu$, $\lambda$ for the boundary of
969: $D_1^2 \times S^1$,
970: $\mu$ for the boundary of $D_2^2 \times S^1$ and $\nu$ for the the boundary of
971: $D^2 \times S^1$, by $Z_\Delta(D^2 \times S^1 \setminus ({\rm Int}D_1^2 \times
972: S^1 \cup {\rm Int}D_2^2 \times S^1); \lambda,\mu;\nu)$.
973: See Figure \ref{solid-b}.
974:
975: \begin{figure}[htbp]
976: \begin{center}
977: \scalebox{0.4}{
978: \includegraphics{solid2.eps}}
979: \end{center}
980: \caption{The labeled 3-manifold obtained by removing two solid tori
981: from a solid torus}
982: \label{solid-b}
983: \end{figure}
984: \bigskip
985:
986: Let $\{ \pi_0, \cdots, \pi_r \}$ be the minimal central projections of
987: ${\rm Tube} \Delta$. Then, we have the following lemma.
988: \begin{lem}\label{innerproduct}
989: $\langle \pi_i, \pi_j \rangle_{Z(S^1 \times S^1)} = \delta_{i,j} n_i^2$
990: ($i,j=0, \dots, r$), where $\langle \cdot, \cdot \rangle_{Z(S^1 \times S^1)}$ is
991: the inner product of $Z_\Delta(S^1 \times S^1)$ defined by the Turaev-Viro-Ocneanu
992: TQFT.
993: \end{lem}
994:
995: \noindent
996: Proof. First, we consider the case $i=j$. We note that $\langle \pi_i, \pi_i
997: \rangle_{Z(S^1 \times S^1)}=Z_\Delta(D^2 \times S^1 \setminus {\rm Int}D_0^2
998: \times S^1;\pi_i, \pi_i)$.
999:
1000: Cut $D^2 \times S^1 \setminus {\rm Int}D_0^2 \times S^1$ along the meridian,
1001: then $T ={\rm annulus} \times [0,1]$ is created. Let $A_{jk}$ be the value of
1002: $Z_\Delta(T, \xi_j, \xi_k$), where $T$ is labeled by $\xi_j$ on one side of
1003: the sections and by $\xi_k$ on the other side. Then, $Z_\Delta(D^2 \times S^1
1004: \setminus {\rm Int}D_0^2 \times S^1;\pi_i, \pi_i)=\sum_{\xi_j} A_{jj}$.
1005: There exists an operator $A$ such that $A_{jk}=\langle A
1006: \xi_j, \xi_k \rangle$, where $\langle \cdot, \cdot \rangle$ is the inner
1007: product on intertwiners on the sections of $T$ so that $\langle \xi_j, \xi_k
1008: \rangle=\delta_{jk}$.
1009:
1010: It is easy to see that the operator $A$ is a projection. Hence,
1011: $\langle \pi_i, \pi_i \rangle_{Z(S^1 \times S^1)}=\sum_{j=1}^m A_{jj} =
1012: {\rm Tr}(A) \in {\mathbb N}$. Namely, we have proved that
1013: $\langle \pi_i, \pi_i \rangle_{Z(S^1 \times S^1)}={\rm dim}\; \pi_i({\rm Tube}
1014: \Delta) = n_i^2$.
1015:
1016: When $i \ne j$, it is easy to see that $\langle \pi_i, \pi_j
1017: \rangle_{Z(S^1 \times S^1)}=0$, because $\pi_i$ and $\pi_j$ are central
1018: projections orthogonal to each other. This ends the proof. \hfill Q.E.D.
1019:
1020: \begin{thm}\label{centerTube}
1021: Let $\Delta$ be a finite system. Then,
1022: the center of ${\rm Tube}\Delta$ is naturally isomorphic to
1023: $Z_\Delta(S^1 \times S^1)$ as a vector space.
1024: \end{thm}
1025:
1026: \noindent
1027: Proof. Let $V(S^1 \times S^1)$ be the linear span of elements in ${\rm Tube}\; \Delta$ such
1028: that the left and right labels in Figure \ref{tube} are equal. Since ${\rm Tube}\Delta
1029: \supset V(S^1 \times S^1)$ and $V(S^1 \times S^1)$ contains the center of
1030: ${\rm Tube}\Delta$, it is enough to consider $V(S^1 \times S^1)$ instead of whole
1031: tube algebra.
1032:
1033: For $X \in V(S^1 \times S^1) $, we set $\varphi_i(X)=Z_\Delta(D^2 \times S^1 \setminus
1034: {\rm Int}D_0^2 \times S^1; \pi_i, X)$. If $X \perp V(S^1 \times S^1)$
1035: with respect to the colored inner product of ${\rm Tube} \Delta$, we set
1036: $\varphi_i(X)=0$. Here, for $X=X_2 X_1^* \in (\xi \cdot \zeta, \zeta \cdot \eta)$, $X_1 \in
1037: (\delta, \xi \cdot \zeta)$, $X_2 \in (\delta, \zeta \cdot \eta)$ and $Y=Y_2 Y_1^* \in
1038: (\xi' \cdot \zeta', \zeta' \cdot \eta')$, $Y_1 \in (\delta', \xi' \cdot \zeta')$,
1039: $Y_2 \in (\delta', \zeta' \cdot \eta')$, the colored inner product
1040: $\langle X, Y \rangle_{\rm color}$ of ${\rm Tube} \Delta$ is defined by
1041: $\delta_{\xi,\xi'}\delta_{\eta,\eta'}\delta_{\zeta,\zeta'} \langle X_1, Y_1 \rangle
1042: \langle X_2, Y_2 \rangle$. The last two brackets stand for the inner
1043: products of the intertwiner spaces $(\delta, \xi \cdot \zeta)$ and
1044: $(\delta, \zeta \cdot \eta)$, respectively. Then, this $\varphi_i$ is a linear functional
1045: defined on ${\rm Tube}\Delta$. It is easy to see that $\varphi_i$ is tracial.
1046:
1047: If $X$ is an element in $\pi_i({\rm Tube}\Delta)$, then we have
1048: \[
1049: \varphi_i(X)=\langle \pi_i, \pi_i \rangle_{Z(S^1 \times S^1)} {\rm tr}(X),
1050: \]
1051: where ${\rm tr}$ is the normalized trace. In a similar way, for $X,Y \in
1052: \pi_i({\rm Tube}\Delta)$, the value
1053: $Z_\Delta(D^2 \times S^1 \setminus {\rm Int}D_0^2 \times S^1; Y, X)$ is
1054: equal to
1055: $\langle \pi_i, \pi_i \rangle_{Z(S^1 \times S^1)} {\rm tr}(X){\rm tr}(Y^*)$.
1056:
1057: When we write $X_j=\pi_jX$ for $X \in V(S^1 \times S^1)$, $X=\oplus_{j=0}^r X_j$.
1058: For $X \in {\rm Tube}\Delta$, we put
1059: \[
1060: E(X)= \sum_{i=0}^r \frac{\varphi_i(X)}{\langle \pi_i, \pi_i
1061: \rangle_{Z(S^1 \times S^1)} } \pi_i=
1062: \sum_{i=0}^r {\rm tr}(X_i) \pi_i.
1063: \]
1064: Then, this is a conditional expectation from ${\rm Tube}\Delta$ to the
1065: center of ${\rm Tube}\Delta$. We have a description of the kernel of $E$ by
1066: ${\rm Ker}E = {\rm the \; linear \; span \; of \; }
1067: \{ X \in V(S^1 \times S^1)|{\rm tr}(X_j)=0, j=0, \dots, r
1068: \} \cup \{ X \in {\rm Tube}\Delta | \langle X, Y \rangle_{\rm color}=0 \; {\rm for \
1069: all}\; Y \in V(S^1 \times S^1) \}$.
1070:
1071: We have
1072: \begin{eqnarray*}
1073: \langle X_j, X_j \rangle_{Z(S^1 \times S^1)}
1074: &=&\langle \pi_j, \pi_j
1075: \rangle_{Z(S^1 \times S^1)} {\rm tr}(X_j){\rm tr}(X_j^*) \\
1076: &=& \langle \pi_j, \pi_j \rangle_{Z(S^1 \times S^1)} |{\rm tr}(X_j)|^2,
1077: \end{eqnarray*}
1078: and we put $Q=\{ X \in V(S^1 \times S^1)|\langle X, X \rangle_{Z(S^1 \times
1079: S^1)}=0 \}$.
1080: Then,
1081: $Z(S^1 \times S^1)=V(S^1 \times S^1)/Q = {\rm Tube}\Delta /{\rm Ker}E
1082: \cong {\rm Center}({\rm Tube}\Delta)$.
1083: Precisely, denoting the embedding map from $Z_\Delta(S^1 \times S^1)$ into
1084: Tube $\Delta$ by $\iota$, we have proved that Center(Tube $\Delta$)
1085: $=\iota(Z_\Delta(S^1 \times S^1)) \subset {\rm Tube} \; \Delta$.
1086: \hfill Q.E.D.
1087:
1088: \begin{rem}
1089: From this theorem and Lemma \ref{innerproduct},
1090: $\{ \frac{\pi_i}{n_i} \}_{i=0}^r$ is an orthonormal basis of
1091: $Z(S^1 \times S^1)$.
1092: \end{rem}
1093:
1094:
1095: \subsection{Verlinde basis}
1096:
1097: In this subsection, we show the existence of a basis of
1098: $Z_\Delta(S^1 \times S^1)$ nicely behaved under the action of
1099: $SL_2(\mathbb{Z})$, which we call the {\it Verlinde basis}. It
1100: deeply depends on the structure of the tube algebra.
1101:
1102: Before the proof of the existence of such basis, we need some preparations.
1103:
1104: Let $p_i$ be a minimal projection in $\pi_i ({\rm Tube} \Delta)$ for
1105: each $i=0,\dots,r$. By Lemma \ref{innerproduct}, we have
1106: $\langle p_i, p_j \rangle_{Z(S^1 \times S^1)} = \delta_{ij}$.
1107: (We use the same notation $p_i$ as an element of $Z_\Delta(S^1 \times S^1)$.)
1108: Then, by the last remark in the previous subsection, we have
1109: \begin{equation}\label{p_i-pi_i}
1110: p_i=\frac{\pi_i}{n_i}\ (i=0,\dots, r)
1111: \end{equation}
1112: in $Z_\Delta(S^1 \times S^1)$.
1113:
1114: Let us compute the value $Z_\Delta(D^2 \times S^1; p_i)$. For this,
1115: we look at $Z_\Delta(D^2 \times S^1; \pi_i)$. It is a summation of
1116: $Z_\Delta(D^2 \times [0,1], \xi_k,\xi_k)$ over the intertwiners $\xi_k$'s
1117: since $D^2 \times [0,1]$ is created by cutting the solid torus along the
1118: meridian, This is nothing but the summation of the dimension
1119: of the partially labeled disks labeled by $p_{i_j}$ over $j$, $j=1,\cdots,
1120: n_i$, where $p_{i_j}$'s are minimal projections such that $\sum_j p_{i_j}=
1121: \pi_i$. (See \cite[Chapter 12]{EK} for the definition of the partially labeled
1122: surfaces.) See Figure \ref{label-pi}.
1123:
1124: \begin{figure}[htbp]
1125: \begin{center}
1126: \scalebox{0.4}{
1127: \includegraphics{label-pi.eps}}
1128: \end{center}
1129: \caption{The decomposition of a labeled disk}
1130: \label{label-pi}
1131: \end{figure}
1132:
1133: When we denote the dimension of the partially labeled disk with label
1134: $p_{i_j}$ by $c_i$, the dimension of the partially labeled disk with label
1135: $\pi_i$ becomes $n_i c_i$.
1136:
1137: Take a summation of $n_i c_i$ over all $i$'s, then it is equal to the
1138: dimension of the partially labeled disk with label $\sum_i \pi_i=1$. Hence,
1139: it is equal to the dimension of the triangulated disk in Figure
1140: \ref{solid-d}, which edge $AB$ is glued together.
1141: (The boudary element is a direct sum taken over arbitray $\rho$.)
1142:
1143: \begin{figure}[htbp]
1144: \begin{center}
1145: \scalebox{0.4}{
1146: \includegraphics{solid4.eps}}
1147: \end{center}
1148: \caption{A triangulated disk}
1149: \label{solid-d}
1150: \end{figure}
1151:
1152: Then, this vector space is non-trivial only in the case of $\rho=id$ and then,
1153: the dimension is one. It means that $\sum_i n_i c_i =1$. Hence,
1154: only one summand can survive. We may and do assume $n_0 c_0=1$. (Hence,
1155: $n_0=1$, $c_0=1$.) In a similar manner, we denote $\pi_i$ by $\pi_0$ in this
1156: case.
1157:
1158: Let us summerize the above argument as a lemma.
1159:
1160: \begin{lem}\label{value-solid}
1161: $Z_\Delta(D^2 \times S^1 \setminus {\rm Int} D_0^2 \times S^1;\pi_i)
1162: =Z_\Delta(D^2 \times S^1 \setminus {\rm Int} D_0^2 \times S^1;p_i)
1163: =\delta_{i0}$.
1164: \end{lem}
1165:
1166: Let $N_{ij}^k$ be $Z_\Delta (D^2 \times S^1 \setminus ( {\rm Int}D_1^2 \times
1167: S^1 \cup {\rm Int}D_2^2 \times S^1);p_i,p_j;p_k )$. See Figure \ref{solid-b}.
1168: Cut $D^2 \times S^1 \setminus ( {\rm Int}D_1^2 \times S^1 \cup
1169: {\rm Int}D_2^2 \times S^1)$ along the meridian, then we have $P=(D^2 \setminus
1170: {\rm Int} D_1^2 \cup {\rm Int} D_2^2) \times [0,1]$, and as a value, $N_{ij}^k$ is
1171: a summation of $Z_\Delta(P, \xi_k, \xi_k)$ over $\xi_k$'s, where $P$ is labeled
1172: by $\xi_k$'s on the sections. This value can be written
1173: $\langle A \xi_k, \xi_k \rangle$, where $\langle \cdot, \cdot \rangle$ is the inner
1174: product of the Hilbert space of the section. Then, it is easy to see that $A$ is a
1175: projection. Hence, $N_{ij}^k={\rm Tr}(A) = {\rm dim } \; H_{\rm pants} \in {\mathbb N}$,
1176: where $H_{\rm pants}$ is the Hilbert space associated with the 3-holed sphere in the
1177: Turaev-Viro-Ocneanu TQFT. See Figure \ref{Nijk}.
1178:
1179: \begin{figure}[htbp]
1180: \begin{center}
1181: \scalebox{0.4}{
1182: \includegraphics{Nijk.eps}}
1183: \end{center}
1184: \caption{$H_{\rm pants}$}
1185: \label{Nijk}
1186: \end{figure}
1187:
1188: With these settings, we can prove the following theorem, which is one of
1189: our main theorems in this paper. (Many of the contents have already appeared
1190: in \cite{EK}, \cite{EK1} following several presentations of Ocneanu, but
1191: some normalizations are missing or incorrect there, so we include a
1192: complete proof here.)
1193: \begin{thm}\label{verlinde-basis}
1194: Let $\Delta$ be a finite system and $\{ p_i \}_{i=0}^r$ minimal
1195: projections such that each $p_i$ belongs to $\pi_i ({\rm Tube}(\Delta))$,
1196: $i=0,\dots, r$. Then,
1197: $\{ p_i \}_{i=0}^r$ is a Verlinde basis of $Z_\Delta(S^1 \times S^1)$
1198: in the sense of Section 2.
1199: \end{thm}
1200:
1201: \noindent
1202: Proof. We check all the conditions for $\{ p_0, \dots, p_r \}$ to be a
1203: Verlinde basis step by step.
1204:
1205: Condition (ii.c.1):
1206: Look at the inner product $n_j^2=\langle \pi_j, \pi_j
1207: \rangle_{Z(S^1 \times S^1)}$. This value is defined by
1208: $Z_\Delta(D^2 \times S^1 \setminus {\rm Int}D_0^2 \times S^1;\pi_j,\pi_j)$.
1209: Cut $D^2 \times S^1 \setminus {\rm Int}D_0^2 \times S^1$ along the meridian,
1210: then we have $X_0={\rm annulus} \times [0,1]$ as topological object.
1211: Since $\pi_j$ is central, $X_0$ can be depicted as in Figure \ref{solid-e},
1212: where $\xi_k$'s are intertwiners on the sections.
1213:
1214: \begin{figure}[htbp]
1215: \begin{center}
1216: \scalebox{0.8}{
1217: \includegraphics{solid5.eps}}
1218: \end{center}
1219: \caption{$X_0$ after changes}
1220: \label{solid-e}
1221: \end{figure}
1222:
1223: Let $X$ be the solid torus that we had at last in Figure \ref{solid-e}.
1224: Then, we have
1225: \begin{equation}
1226: \langle \pi_j, \pi_j \rangle_{Z(S^1 \times S^1)}=\sum_{\xi_k} \alpha(\xi_k)
1227: Z_\Delta (X),
1228: \end{equation}
1229: where $\alpha(\xi_k)'s$ are positive coefficients determined by
1230: some products of statistical dimensions.
1231: It is obvious that each $Z_\Delta (X)$ is non-negative.
1232:
1233: Since $S(p_j)=\sum_i S_{ij} p_i$, it follows from Lemma \ref{value-solid} that
1234: $Z_\Delta (D^2\times S^1;S(p_j))= S_{0j}$.
1235:
1236: It is now easy to see that the value of the solid torus labeled by
1237: $S(p_j)$ has a similar expression of the summation when it is cut along the
1238: meridian. Namely, $S_{0j}$ is written in the form $S_{0j}=\sum_{\xi_k}
1239: \beta(\xi_k) Z_\Delta (X)$, where $\beta(\xi_k)$'s are strictly positive
1240: coefficients detemined by $n_j$ and some products of statistical dimensions.
1241: As we saw, not all of the values of $Z_\Delta (X)$ can be zero, so
1242: one of them must be strictly positive. Hence, $S_{0j} > 0$.
1243:
1244: Condition (ii.a), (iii): Since both $S$ and $T$-matrices are
1245: unitary, we check that $S$ is symmetric and $T$ is diagonal with repsect to
1246: $\{p_0, \dots, p_r \}$.
1247:
1248: First we prove that $S$ is symmetric. From a topological observation of
1249: the $Z(S)$-action on $Z(S^1 \times S^1)$,
1250: we have
1251: \begin{eqnarray*}
1252: S_{ij}
1253: &=&\langle S(p_i), p_j \rangle_{Z(S^1 \times S^1)}
1254: = \overline{\langle S(p_i)^*, p_j^* \rangle}_{Z(S^1 \times S^1)}
1255: = \overline{\langle S(p_i)^*, p_j \rangle}_{Z(S^1 \times S^1)} \\
1256: &=& \overline{\langle S^*(p_i), p_j \rangle}_{Z(S^1 \times S^1)}
1257: = \overline{\langle p_i, S(p_j) \rangle}_{Z(S^1 \times S^1)}
1258: = \langle S(p_j), p_i \rangle_{Z(S^1 \times S^1)} \\
1259: &=&S_{ji}.
1260: \end{eqnarray*}
1261:
1262: Next, we prove that $T$ is diagonal. From a topological observation of
1263: the $Z(T)$-action on $Z(S^1 \times S^1)$, we have
1264: \[
1265: \langle T(X) \cdot T^*(Y), Z \rangle_{Z(S^1 \times S^1)} =
1266: \langle X \cdot Y, Z \rangle_{Z(S^1 \times S^1)}
1267: \]
1268: for any $X,Y,Z \in V(S^1 \times S^1)$, where $\cdot$ stands for the multiplication
1269: in the tube algebra. Hence, we have $T(X) \cdot T^*(Y)=X \cdot Y$ for any
1270: $X, Y \in V(S^1 \times S^1)$. This implies that $T$ is diagonal.
1271:
1272: Condition (ii.c.2): (Verlinde identity) From the definition of the fusion
1273: algebra associated with the Turaev-Viro-Ocneanu TQFT, $N_{ij}^k$ is a structure
1274: constant of the fusion algebra.
1275:
1276: From the unitarity of $S$, it is enough to prove
1277: \begin{equation}\label{verlinde-id}
1278: \sum_{i,j} N_{ij}^k \; \overline{S_{im}}\; \overline{S_{jn}} = \delta_{mn}
1279: \frac{\overline{S_{mk}}}{S_{m0}}.
1280: \end{equation}
1281: The first term of the left-hand side of the equation (\ref{verlinde-id}) is
1282: written as $Z_\Delta (D^2 \times S^1 \setminus ({\rm Int}D_1^2
1283: \times S^1 \cup {\rm Int}D_2^2 \times S^1); p_i,p_j;p_k)$. The second and
1284: the third terms of the left-hand side of the equation (\ref{verlinde-id}) can be
1285: written as $Z_\Delta (D^2 \times S^1 \setminus {\rm Int }D_0^2 \times S^1;
1286: p_i, S(p_m))$, $Z_\Delta (D^2 \times S^1 \setminus {\rm Int}D_0^2 \times S^1;
1287: p_j, S(p_n))$, respectively.
1288: Use gluing and embed the second and third terms into the first term, then
1289: the left-hand side of the equation (\ref{verlinde-id}) is equal to
1290: $Z_\Delta (D^2 \times S^1 \setminus ({\rm Int}D_1^2 \times S^1 \cup
1291: {\rm Int}D_2^2 \times S^1); S(p_m),S(p_n);p_k)$. Then, cut $D^2 \times S^1
1292: \setminus ({\rm Int}D_1^2 \times S^1 \cup {\rm Int}D_2^2 \times S^1$ along
1293: the meridian.
1294:
1295: \begin{figure}[htbp]
1296: \begin{center}
1297: \scalebox{0.45}{
1298: \includegraphics{pi_m-pi_n.eps}}
1299: \end{center}
1300: \caption{}
1301: \label{pi_m-pi_n}
1302: \end{figure}
1303:
1304: By using the gluing axiom of TQFT, we separate the rectangular part of
1305: this cut object by taking the summation of $Z_\Delta$ of the 3-manifold
1306: depicted in Figure \ref{pi_m-pi_n} over orthonormal basis $\eta_l$'s
1307: on the rectangle. Then, by the identity (\ref{p_i-pi_i}), we get the value 0
1308: if $m \ne n$ because $\pi_m$ and $\pi_n$ are central orthogonal projections.
1309: Hence,
1310: \begin{eqnarray*}
1311: &&Z_\Delta (D^2 \times S^1 \setminus ({\rm Int}D_1^2 \times S^1 \cup
1312: {\rm Int}D_2^2 \times S^1); S(p_m),S(p_n);p_k) \\
1313: &=& \delta_{mn} \frac{\sum_m Z_\Delta (D^2 \times S^1 \setminus ({\rm Int}D_1^2
1314: \times S^1 \cup{\rm Int} D_2^2 \times S^1); S(p_m),S(p_n);p_k) \;
1315: Z_\Delta (D^2 \times S^1;S(p_m))}{Z_\Delta (D^2 \times S^1;S(p_n))} \\
1316: &=& \delta_{mn} \frac{Z_\Delta (D^2 \times S^1 \setminus {\rm Int}D_0^2 \times
1317: S^1;S(p_m), p_k)}{Z_\Delta (D^2 \times S^1;S(p_n))} \\
1318: &=& \delta_{mn} \frac{\overline{S_{mk}}}{S_{n0}}.
1319: \end{eqnarray*}
1320:
1321: Condition (i): From the Verlinde identity and the unitarity of $S$, we have
1322: \[
1323: N_{0j}^k=\sum_{l=0}^r \frac{S_{0l} S_{jl} \overline{S_{kl}}}{S_{0l}}
1324: = \sum_{l=0}^r S_{jl} \overline{S_{kl}} = \delta_{jk}.
1325: \]
1326: So $p_0$ is the identitiy element in the fusion algebra.
1327:
1328: Condition (ii.b): Since $S^2$ is $*$-antiisomorphism, it is clear that
1329: $S^2(p_i)$'s are minimal projections again.
1330:
1331: We have $S^2(p_0)=p_0$ because the definition of $p_0$ is invariant
1332: under the 180 degree rotation.
1333:
1334: Condition (iv): We note that $Z(U)$ is nothing but the $*$-operation of
1335: ${\rm Tube}\; \Delta$. So $Z(U)(p_i)=p_i^*=p_i$. The canonical map $\theta$ maps
1336: an orthonormal basis in $Z(-S^1 \times S^1)$ to the dual basis in $Z(S^1
1337: \times S^1)^*$. So, $\theta(p_i)=\widehat{p_i}$, where
1338: $\{ \widehat{p_i} \}_{i=0}^r$ is a dual basis of $\{ p_i \}_{i=0}^r$ in
1339: $Z(S^1 \times S^1)^*$ such that $\widehat{p_i}(p_j)=\delta_{ij}$. Getting
1340: together, we have $\theta \circ Z(U) (p_i) = \theta(p_i)=\widehat{p_i}$.
1341: \hfill Q.E.D.
1342:
1343:
1344:
1345: \section{Applications}
1346:
1347: From the conclusion of Section 4, we know that there exists a Verlinde basis
1348: of $Z_{\Delta}(S^1 \times S^1)$ in the sense of Section 2 in a
1349: Turaev-Viro-Ocneanu TQFT. Hence, for a closed 3-manifold $M$, we have
1350: the following Dehn surgery formula:
1351: \[
1352: Z_\Delta (M) = \sum_{i_1, \dots, i_m=0}^r S_{0i_1} \cdots S_{0i_m}
1353: J(L;i_1, \dots, i_m),
1354: \]
1355: where we have assumed that the manifold $M$ is obtained from $S^3$ by Dehn
1356: surgery along a framed link $L=L_1 \cup \cdots \cup L_m$.
1357:
1358: The purpose in this section is to understand the formula of
1359: right-hand side of the above equation by introducing a notion of
1360: the {\it tube system} due to Ocneanu \cite{Ocneanu,O}, which gives a tensor
1361: category.
1362:
1363: \subsection{Tube systems}
1364:
1365: Let us start with the definition of a tube system. Let $\Delta$ be a
1366: finite system. A {\it tube system} ${\cal D}(\Delta)$ is a tensor category
1367: defined in the following way.
1368: First of all, the objects of ${\cal D}(\Delta)$ are the ${\mathbb C}$-linear
1369: span of all minimal projections in the tube algebra Tube\/$\Delta$.
1370: For minimal projections $p_i$ and $p_j$, the hom-set ${\rm Hom}(p_i, p_j)$
1371: is the set of vectors in the labeled surface depicted in Figure \ref{p_i-p_j}.
1372: For general objects $X, Y \in {\cal D}(\Delta)$, we define ${\rm Hom} (X,Y)$ by
1373: extending ${\rm Hom}(p_i, p_j)$ by linearity.
1374: \begin{figure}[htbp]
1375: \begin{center}
1376: \scalebox{0.47}{
1377: \includegraphics{p_i-p_j.eps}}
1378: \end{center}
1379: \caption{Hom$(p_i,p_j)$}
1380: \label{p_i-p_j}
1381: \end{figure}
1382: Let $\sim$ be the
1383: Murray-von Neumann equivalence relation between projections. (Namely,
1384: two projections $p \; ,q \in {\rm Tube}\; \Delta$ are equivalent in the sense
1385: of Murray-von Neumann if there exists a partial isometry $v \in {\rm Tube} \;
1386: \Delta$ such that $p=v^* v$ and $q=v v^*$.) If $p_i \sim p_j$, then
1387: ${\rm Hom}(p_i, p_j) \cong \mathbb{C}$ and if $p_i \nsim p_j$, then
1388: ${\rm Hom}(p_i, p_j) =\{ 0 \}$. Hence, minimal projections are
1389: simple objects in ${\cal D}(\Delta)$.
1390: We denote ${\rm Hom}(p_i, p_i)$ by ${\rm End}(p_i)$ for simplicity.
1391:
1392: For simple objects $p,q$ and $r$, the composition $x \cdot y$ of $x \in
1393: {\rm Hom} (p,q)$ and $y \in {\rm Hom}(q,r)$ is defined by the concateneation
1394: of $x$ followed by $y$. See Figure \ref{composition}. For general objects $p,q$
1395: and $r$, we define the composition of two morphisms as above by linearlity.
1396: We also define $x^*$ for $x$ in ${\rm Hom}(p,q)$ by inversing $x$ inside
1397: out.
1398: \begin{figure}[htbp]
1399: \begin{center}
1400: \scalebox{0.4}{
1401: \includegraphics{composition.eps}}
1402: \end{center}
1403: \caption{The composition of $x \in {\rm Hom} (q,r)$ and $y \in {\rm Hom}(p,q)$}
1404: \label{composition}
1405: \end{figure}
1406:
1407: Let $p_j$ be a simple object and $q$ be an object in ${\cal D}(\Delta)$. We
1408: define an inner product $\langle x, y \rangle$ of $x, y \in {\rm Hom}(q, p_i)$ by
1409: $\langle x, y \rangle= x \cdot y^* $, as a composition of morphisms. Then, with this
1410: inner product, ${\rm Hom}(q, p_i)$ becomes a Hilbert space.
1411:
1412: Next, we define the tensor product $p_i \otimes p_j$ of two simple objects
1413: $p_i$ and $p_j$ by the fusion product $p_i * p_j$ which was defined in
1414: Section 2.
1415: Then, ${\rm Hom} (p_i \otimes p_j, p_k)$ consists of vectors in the labeled
1416: surface of Figure \ref{Nijk}. We define an inner product $\langle x, y
1417: \rangle$ of $x, y \in {\rm Hom} (p_i \otimes p_j, p_k)$ by a composition of two
1418: morphisms $\langle x, y \rangle= x \cdot y^* \in {\rm End}(p_k) \cong \mathbb{C}$.
1419: (Here, $y^*$ is defined by inversing $y$ (pants) inside out.) Note that we have
1420: Frobenius reciprocities for morphisms of ${\rm Hom}(p_i \otimes p_j, p_k)$.
1421: For instance, we obtain ${\rm Hom}(p_i \otimes p_j, p_k) \cong {\rm Hom}(
1422: p_i, p_k \otimes p_{\bar{j}})$. See Figure \ref{frobenius}. Frobenius
1423: reciprocities are given by the graphical operations in the category
1424: ${\cal D}(\Delta)$.
1425: \begin{figure}[htbp]
1426: \begin{center}
1427: \scalebox{0.65}{
1428: \includegraphics{frobenius.eps}}
1429: \end{center}
1430: \caption{A Frobenius reciprocity map}
1431: \label{frobenius}
1432: \end{figure}
1433:
1434: We call the above defined semi-simple tensor category ${\cal D}(\Delta)$
1435: a {\it tube system}.
1436:
1437: To compute the quantum dimension of $p_i$, we make the composition of
1438: morphisms $b_i^* \circ b_i$, where $b_i$ is the distinguished morphism
1439: from $p_0$ to $p_i \otimes p_{\bar{i}}$ obtained from Frobenius
1440: reciprocity of $id_{p_i} \in {\rm End}(p_i)$. It must be a scalar multiple
1441: of $p_0$ since $p_0$ is a simple object. We put this value $c$.
1442: (See Figure \ref{qdim}.)
1443: \begin{figure}[htbp]
1444: \begin{center}
1445: \scalebox{0.6}{
1446: \includegraphics{q-dim.eps}}
1447: \end{center}
1448: \caption{Computing a quantum dimension}
1449: \label{qdim}
1450: \end{figure}
1451:
1452: To obtain the value $c$, we connect the upper $p_0$ and the lower $p_0$,
1453: and embed it into $S^3$. Fill out the outside of the tube, and take the
1454: Turaev-Viro-Ocneanu invariant of it. Then, we get the
1455: $c=J(\text{\LARGE $\bigcirc $}; i)/S_{00}$. This is the quantum dimension of
1456: $p_i$, denoted by ${\rm dim}\; p_i$.
1457:
1458: We define the map $c_{p_i,p_j}$ from $p_i \otimes p_j$ to
1459: $p_j \otimes p_i$ as in Figure \ref{braiding}.
1460: \begin{figure}[htbp]
1461: \begin{center}
1462: \scalebox{0.8}{
1463: \includegraphics{braiding.eps}}
1464: \end{center}
1465: \caption{A map $c_{p_i,p_j}$}
1466: \label{braiding}
1467: \end{figure}
1468: It is easy to see that this map $c_{p_i,p_j}$ satisfies the axioms of
1469: a braiding, since the map $c_{p_i,p_j}$ is defined in a topological
1470: way. So, we now know the tensor category ${\cal D}(\Delta)$ is braided.
1471:
1472: We further define a map $\theta_{p_i} \in {\rm End}(p_i)$ by Figure
1473: \ref{twist}, where $Z_\Delta$ is evaluated at the 3-ball removed a twisted
1474: solid tube and the two tubes in the boudaries of it are both labeled by $p_i$.
1475: \begin{figure}[htbp]
1476: \begin{center}
1477: \scalebox{0.8}{
1478: \includegraphics{twist-tube.eps}}
1479: \end{center}
1480: \caption{The twist map}
1481: \label{twist}
1482: \end{figure}
1483: This map satisfies the following two equalities. (By the definition of
1484: $\theta_{p_i}$, these equalities are proven by making pictures corresponding
1485: to the formulas in both sides and using some topological moves.)
1486: \begin{eqnarray*}
1487: & \theta_{p_i \otimes p_j} = c_{p_j,p_i} \circ c_{p_i,p_j} \circ (\theta_{p_i}
1488: \otimes \theta_{p_j}), \\
1489: & (\theta_{p_i} \otimes id_{p_{\bar{i}}}) \circ b_i = (id_{p_i} \otimes
1490: \theta_{p_{\bar{i}}} ) \circ b_i.
1491: \end{eqnarray*}
1492: Namely, $\theta_{p_i}$ defines a twist on ${\cal D}(\Delta)$ and this makes
1493: ${\cal D}(\Delta)$ a ribbon category.
1494:
1495: Let us make the following compositions of morphisms in ${\cal D}(\Delta)$.
1496: \begin{eqnarray*}
1497: &b_0&
1498: \circ (id_{p_0} \otimes b_j)
1499: \circ (b_i \otimes id_{p_j} \otimes id_{b_{\bar{j}}})
1500: \circ (id_{p_i} \otimes c_{p_{\bar{i}},p_{j}} \otimes id_{p_{\bar{j}}})
1501: \circ (id_{p_i} \otimes c_{p_j, p_{\bar{i}}} \otimes id_{p_{\bar{j}}} ) \\
1502: &\circ& (b_i^* \otimes id_{p_j} \otimes p_{\bar{j}})
1503: \circ (id_{p_0} \otimes b_j^*)
1504: \circ b_0^*
1505: \end{eqnarray*}
1506: Then, it is a scalar multiple of $p_0$ and makes a Hopf link $H$ as a
1507: diagram. (See Figure \ref{hopf}.)
1508: \begin{figure}[htbp]
1509: \begin{center}
1510: \scalebox{0.4}{
1511: \includegraphics{Sij.eps}}
1512: \end{center}
1513: \caption{The Hopf link}
1514: \label{hopf}
1515: \end{figure}
1516: Let us denote this scalar of the Hopf link $H$ by $s_{ij}$. Embed this
1517: compositions of tubes into $S^3$ and
1518: fill out the outside of tubes. By Proposition \ref{Proposition2} in Section 2,
1519: this provides
1520: \[
1521: s_{ij}=\frac{J(H; i, j)}{S_{00}}=\frac{S_{ij}}{S_{00}},
1522: \]
1523: where $S=(S_{ij})_{i,j=0}^r$ is the $S$-matrix with respect to our Verlinde basis
1524: $\{ p_0, \dots, p_r \}$.
1525: Since the $S$-matrix is unitary, the matrix $(s_{ij})_{i,j=0}^r$ is
1526: invertible. It means that our category ${\cal D}(\Delta)$ is modular.
1527:
1528: \begin{rem}
1529: The notions of the tube algebra and the tube system are also described by
1530: the language of a II$_1$-subfactor $N \subset M$ with finite Jones index
1531: and finite depth, although our exposition here uses an infinite
1532: subfactor.
1533: \end{rem}
1534:
1535: \subsection{Dehn surgery formula as a Reshetikhin-Turaev invariant}
1536:
1537: Let $\cal C$ be a modular category and $\{ V_i \}_{i=0}^r$ its simple
1538: objects. Put $\Delta=\Delta_+ = \sum_{i=0}^r t_i^{-1} ({\rm dim}\; V_i)^2$,
1539: $\Delta_- = \sum_{i=0}^r t_i ({\rm dim}\; V_i)^2$ and
1540: $D=(\sum_{i=0}^r ({\rm dim} V_i)^2)^{1/2}$. (Here, for $\Delta$, we
1541: followed the notation in \cite{Turaev}.)
1542: Let $M$ be a closed 3-manifold obtained from $S^3$ by Dehn surgery along a
1543: framed link $L=L_1 \cup \cdots \cup L_m$. Then, the Reshetikhin-Turaev invariant
1544: of $M$ is given by the formula
1545: \[
1546: \tau(M) = \Delta^{\sigma(L)} D^{-\sigma(L)-m-1} \sum_{\lambda \in Col(L)}
1547: (\prod_{n=1}^m {\rm dim} V_{\lambda(n)}) F(L,\lambda),
1548: \]
1549: where $\sigma(L)$ is the signature of $L$ and $F(L,\lambda)$ is
1550: the invariant of the colored framed link $(L, \lambda)$.
1551: (See \cite{Turaev} for the details. Also see \cite{RT}.)
1552:
1553: We note that in the above formula ${\rm dim} V_{\lambda(n)} =
1554: s_{0\lambda(n)}$. Hence the original Rehshtikhin-Turaev formula can be
1555: rewritten with the $s$-matrix in the form
1556: \[
1557: \tau(M) = \Delta^{\sigma(L)} D^{-\sigma(L)-m-1} \sum_{\lambda \in Col(L)}
1558: (\prod_{n=1}^m s_{0 \lambda(n)}) F(L,\lambda).
1559: \]
1560:
1561: Now, we start with the modular category ${\cal D}(\Delta)$ defined in
1562: Section 5.1 instead of a
1563: general modular category and make the
1564: Rehsetikhin-Turaev formula. In our case, we already have $S$-matrix
1565: $( S_{ij} )_{i,j=0}^r$ from the Turaev-Viro-Ocneanu TQFT, which is expressed
1566: with respect to a Verlinde basis $\{ p_0, \dots, p_r \}$ in
1567: $Z_\Delta(S^1 \times S^1)$, and since we know that
1568: $\dim \; p_i=S_{0i}/S_{00}$, by using $D=1/S_{00}$, we can rewrite the
1569: Reshetikhin-Turaev formula in the form
1570: \[
1571: \tau(M) = \Delta^{\sigma(L)} D^{-\sigma(L)} \sum_{\lambda \in Col(L)}
1572: (\prod_{n=1}^m S_{0 \lambda(n)}) S_{00} F(L,\lambda).
1573: \]
1574:
1575: We will prove that $D=\Delta$.
1576: First of all, from the equality $Z_\Delta (S^3)=Z_\Delta (L(1,1))
1577: =\sum_{i=0}^r t_i^{-1} S_{0i}^2$, we have $\frac{1}{S_{00}}=
1578: \lambda=\sum_{i=0}^r t_i^{-1} S_{0i}^2 $ in our notation. From this, we get
1579: \[
1580: \frac{1}{S_{00}}=\lambda =\sum_{i=0}^r t_i^{-1} (\frac{S_{0i}}{S_{00}})^2
1581: = \sum_{i=0}^r t_i^{-1}({\rm dim} \; p_i)^2=\Delta_+.
1582: \]
1583: Taking the complex conjuegation of the above formula, we get
1584: \[
1585: \lambda = \sum_{i=0}^r t_i ({\rm dim} \; p_i)^2=\Delta_-.
1586: \]
1587: Thus $\lambda^2 =\Delta_+ \Delta_- = \sum_{i=0}^r
1588: ({\rm dim} \; p_i)^2$. Namely, we have $D=\Delta$. So, the Reshetikhin-Turaev
1589: invariant constructed from ${\cal D}(\Delta)$ is given by
1590: \[
1591: \tau(M) = \sum_{\lambda \in Col(L)} (\prod_{n=1}^m S_{0 \lambda(n)})
1592: S_{00} F(L,\lambda).
1593: \]
1594:
1595: Since $F$ is uniquely determined by the category of the ribbon tangles
1596: \cite[Part I, Chapter I]{Turaev}, taking the normalization into consideration,
1597: we have $F(L,\lambda)=\frac{J(L,\lambda)}{S_{00}}$. Namely, in our
1598: case,
1599: \[
1600: \tau(M) = \sum_{\lambda \in Col(L)} (\prod_{n=1}^m S_{0 \lambda(n)})
1601: J(L,\lambda),
1602: \]
1603: which is nothing but our Dehn surgery formula Proposition \ref{Proposition1}. Let us
1604: summerize this argument as a theorem.
1605:
1606: \begin{thm}\label{main}
1607: Let $\Delta$ be a finite system.
1608: For a closed oriented 3-manifold $M$, the Reshetikhin-Turaev invariant $\tau(M)$
1609: constructed from a tube system ${\cal D} (\Delta)$ coincides with the
1610: Turaev-Viro-Ocneanu invariant $Z_\Delta (M)$.
1611: \end{thm}
1612:
1613: The following corollary has been proven by several authors \cite{Turaev},
1614: \cite{Roberts} etc in various settings.
1615:
1616: \begin{cor}\label{split}
1617: Let $\Delta$ be a finite $C^*$-tensor category arising from a subfactor.
1618: If $\Delta$ is a modular category, then ${\cal D} (\Delta)$ is equivalent
1619: to $\Delta \otimes \Delta^{\rm op}$ and this provides us with $\tau(M) \cdot
1620: \overline{\tau(M)}=Z_{\Delta}(M)$ for a closed oriented 3-manifold $M$,
1621: where $\tau$ is the Reshetikhin-Turaev invariant constructed from $\Delta$.
1622: \end{cor}
1623:
1624: \noindent
1625: Proof. In \cite{EK2}, it is proved that ${\cal D}(\Delta)$ is equivalent
1626: to $\Delta \otimes \Delta^{\rm op}$ when $\Delta$ is a modular category.
1627: (See Appendix too.) Hence, the rest is clear from Theorem \ref{main}.
1628: \hfill Q.E.D.
1629:
1630: \appendix
1631:
1632:
1633: \section{Appendix}
1634: In this Appendix, we fix inaccuracies in \cite{EK} and \cite{EK1}.
1635: In \cite{EK} and \cite{EK1}, the authors have analyzed the structure of
1636: $M_\infty$-$M_\infty$ bimodules obtained from the {\it asymptotic
1637: inclusion} $M \vee M^{\rm op} \subset M_\infty$ starting from
1638: the inclusion of AFD II$_1$ factors $N \subset M$ with finite Jones index
1639: and finite depth, following Ocneanu.
1640:
1641: Let $\{p_0, \cdots, p_r \}$ be a representative set of the equivalence classes of
1642: minimal projections of Tube$\Delta$ by the Murray-von Neumann equivalence relation.
1643:
1644: We present the $M_\infty$-$M_\infty$ bimodule $X(p_i)$ by Figure \ref{a1},
1645: where the circle at the middle is empty and the annulus around it is
1646: labeled with the minimal projection $p_i$.
1647: (See Chapter 12 in \cite{EK} for more explanation of this kind of pictures.)
1648: \unitlength 0.5mm
1649: \thicklines
1650: \begin{figure}[htpb]
1651: \begin{center}\begin{picture}(160,100)
1652: \put(70,50){\circle{20}}
1653: \put(70,50){\circle{40}}
1654: \put(70,60){\circle*{1}}
1655: \put(70,70){\circle*{1}}
1656: \spline(70,70)(59,65)(53,58)(53.5,40.5)(71,32)(84,40)(83,57)(70,60)
1657: \put(72,34){\vector(1,0){0}}
1658: \put(70,67){\vector(0,1){0}}
1659: \put(70,60){\line(0,1){10}}
1660: \put(50,50){\vector(0,-1){0}}
1661: \put(50,45){\vector(0,-1){0}}
1662: \put(60,50){\vector(0,-1){0}}
1663: \put(60,45){\vector(0,-1){0}}
1664: \put(70,50){\ellipse{60}{80}}
1665: \put(70,50){\ellipse{120}{80}}
1666: \put(10,50){\vector(0,-1){0}}
1667: \put(40,50){\vector(0,-1){0}}
1668: \put(100,50){\vector(0,-1){0}}
1669: \put(130,50){\vector(0,-1){0}}
1670: \put(70,10){\vector(1,0){0}}
1671: \put(70,10){\vector(-1,0){0}}
1672: \put(70,10){\circle*{1}}
1673: \put(70,90){\circle*{1}}
1674: \put(11.9,40){\circle*{1}}
1675: \put(11.9,60){\circle*{1}}
1676: \put(128.1,40){\circle*{1}}
1677: \put(128.1,60){\circle*{1}}
1678: \put(18,30){\circle*{1}}
1679: \put(18,70){\circle*{1}}
1680: \put(122,30){\circle*{1}}
1681: \put(122,70){\circle*{1}}
1682: \put(30.3,20){\circle*{1}}
1683: \put(30.3,80){\circle*{1}}
1684: \put(109.7,20){\circle*{1}}
1685: \put(109.7,80){\circle*{1}}
1686: \spline(30.3,80)(40,86)(60,89)(70,90)(80,89)(100,86)(109.7,80)
1687: \spline(18,70)(40,84)(60,89)(70,90)(80,89)(100,84)(122,70)
1688: \spline(11.9,60)(40,84)(60,88)(70,90)(80,88)(100,84)(128.1,60)
1689: \spline(11.9,40)(40,83)(60,87)(70,90)(80,87)(100,83)(128.1,40)
1690: \spline(18,30)(40,80)(60,87)(70,90)(80,87)(100,80)(122,30)
1691: \spline(30.3,20)(40,78)(60,87)(70,90)(80,87)(100,78)(109.7,20)
1692: %\put(70,3){\makebox(0,0){Fig. 4.14}}
1693: \end{picture}
1694: \end{center}
1695: \caption{The bimodule $X(p_i)$}
1696: \label{a1}
1697: \end{figure}
1698: We can prove that each $M_\infty$-$M_\infty$ bimodule $X(p_i)$ $(i=0,
1699: \cdots, r)$ is irreducible in a similar way to the proof of
1700: Theorem 12.26 in \cite{EK}.
1701:
1702: Since $[{}_{M_\infty} X(p_i)_{M_\infty}]^{1/2}=\dim \; p_i$,
1703: we have the equalities $\sum_{i=0}^r [X(p_i)]=\sum_{i=0}^r (\dim \; p_i)^2
1704: =\lambda^2$, where dim $p_i$ is the quantum dimension of $p_i$ as an object of
1705: ${\cal D}(\Delta)$. On the other hand, the global index of the asymptotic
1706: inclusion
1707: is given by $\lambda^2$. This means that all the irreducible
1708: $M_\infty$-$M_\infty$ bimodules obtained from the asymptotic inclusion are
1709: given by $\{X(p_i) \}_{i=0}^r$.
1710:
1711: \begin{rem}
1712: In \cite{EK} and \cite{EK1}, irreducible bimodules are labeled by $\pi_i$'s,
1713: i.e., minimal central projections of Tube$\Delta$, instead of the minimal
1714: projections of it, which is incorrect.
1715: \end{rem}
1716:
1717: We describe the fusion rule of irreducble $M_\infty$-$M_\infty$
1718: bimodules. The relative tensor product of two $M_\infty$-$M_\infty$
1719: bimodules $X(p_i) \otimes_{M_\infty} X(p_j)$ is decomposed into irreducible
1720: bimodules as in Figure \ref{a5}.
1721:
1722: \unitlength 0.45mm
1723: \thicklines
1724: \begin{figure}[htpb]
1725: \begin{center}\begin{picture}(240,100)
1726: \put(170,50){\circle{20}}
1727: \put(170,50){\circle{14}}
1728: \put(45,50){\circle{20}}
1729: \put(45,50){\circle{14}}
1730: \put(75,50){\circle{20}}
1731: \put(75,50){\circle{14}}
1732: \put(60,50){\circle{60}}
1733: \put(60,50){\circle{66}}
1734: \put(160,50){\vector(0,-1){0}}
1735: \put(163,50){\vector(0,-1){0}}
1736: \put(30,50){\vector(0,-1){0}}
1737: \put(27,50){\vector(0,-1){0}}
1738: \put(35,50){\vector(0,-1){0}}
1739: \put(38,50){\vector(0,-1){0}}
1740: \put(65,50){\vector(0,-1){0}}
1741: \put(68,50){\vector(0,-1){0}}
1742: \put(170,50){\ellipse{60}{80}}
1743: \put(170,50){\ellipse{120}{80}}
1744: \put(110,50){\vector(0,-1){0}}
1745: \put(140,50){\vector(0,-1){0}}
1746: \put(200,50){\vector(0,-1){0}}
1747: \put(230,50){\vector(0,-1){0}}
1748: \put(170,10){\vector(1,0){0}}
1749: \put(170,10){\vector(-1,0){0}}
1750: \put(170,10){\circle*{1}}
1751: \put(170,90){\circle*{1}}
1752: \put(111.9,40){\circle*{1}}
1753: \put(111.9,60){\circle*{1}}
1754: \put(228.1,40){\circle*{1}}
1755: \put(228.1,60){\circle*{1}}
1756: \put(118,30){\circle*{1}}
1757: \put(118,70){\circle*{1}}
1758: \put(222,30){\circle*{1}}
1759: \put(222,70){\circle*{1}}
1760: \put(130.3,20){\circle*{1}}
1761: \put(130.3,80){\circle*{1}}
1762: \put(209.7,20){\circle*{1}}
1763: \put(209.7,80){\circle*{1}}
1764: %+100
1765: \spline(130.3,80)(140,86)(160,89)(170,90)(180,89)(200,86)(209.7,80)
1766: \spline(118,70)(140,84)(160,89)(170,90)(180,89)(200,84)(222,70)
1767: \spline(111.9,60)(140,84)(160,88)(170,90)(180,88)(200,84)(228.1,60)
1768: \spline(111.9,40)(140,83)(160,87)(170,90)(180,87)(200,83)(228.1,40)
1769: \spline(118,30)(140,80)(160,87)(170,90)(180,87)(200,80)(222,30)
1770: \spline(130.3,20)(140,78)(160,87)(170,90)(180,87)(200,78)(209.7,20)
1771: \put(170,35){\makebox(0,0){$p_k$}}
1772: \put(45,35){\makebox(0,0){$p_i$}}
1773: \put(75,35){\makebox(0,0){$p_j$}}
1774: \put(60,10){\makebox(0,0){$p_k$}}
1775: \put(12,47){\makebox(0,0){$\displaystyle\sum_k$}}
1776: \put(100,50){\makebox(0,0){$\otimes$}}
1777: %\put(120,3){\makebox(0,0){Fig. 4.15}}
1778: \end{picture}
1779: \end{center}
1780: \caption{Decomposition of a bimodule}
1781: \label{a5}
1782: \end{figure}
1783: Hence, the fusion rule is given by the fusion rule of the fusion algebra
1784: associated with the Turaev-Viro-Ocneanu TQFT.
1785:
1786: It is easy to see that $M_\infty$-$M_\infty$ bimodules arising from the
1787: asymptotic inclusion $M \vee M^{\rm op} \subset M_\infty$ give rise to
1788: a modular category with the same braiding and twist as ones in a tube system.
1789: We denote this category of $M_\infty$-$M_\infty$ bimodules by ${\cal M}_\infty$.
1790: To a simple object $X(p_i)$ in ${\cal M}_\infty$, we assign a simple object
1791: $p_i$ in ${\cal D}(\Delta)$.
1792: From Figure \ref{a5}, it is easy to see that we have a functor
1793: $F: \; {\cal M}_\infty \longrightarrow {\cal D}(\Delta)$, when we look at
1794: the morphisms in ${\cal M}_\infty$.
1795:
1796: Let us now consider the opposite direction to the functor $F$.
1797: For a given morphism $x \in {\rm Hom} (p_i \otimes p_j, p_k)$ of
1798: ${\cal D}(\Delta)$, we will construct a homomoprhism in ${\rm Hom} (X(p_i)
1799: \otimes_{M_\infty} X(p_j), X(p_k))$.
1800:
1801: \unitlength 0.5mm
1802: \thicklines
1803: \begin{figure}[htpb]
1804: \begin{center}\begin{picture}(160,100)
1805: \put(85,50){\circle{20}}
1806: \put(55,50){\circle{20}}
1807: \put(85,50){\circle{14}}
1808: \put(55,50){\circle{14}}
1809: \put(45,50){\vector(0,-1){0}}
1810: \put(75,50){\vector(0,-1){0}}
1811: \put(48,50){\vector(0,-1){0}}
1812: \put(78,50){\vector(0,-1){0}}
1813: \put(70,50){\ellipse{60}{80}}
1814: \put(70,50){\ellipse{120}{80}}
1815: \put(10,50){\vector(0,-1){0}}
1816: \put(40,50){\vector(0,-1){0}}
1817: \put(100,50){\vector(0,-1){0}}
1818: \put(130,50){\vector(0,-1){0}}
1819: \put(70,10){\vector(1,0){0}}
1820: \put(70,10){\vector(-1,0){0}}
1821: \put(70,10){\circle*{1}}
1822: \put(70,90){\circle*{1}}
1823: \put(11.9,40){\circle*{1}}
1824: \put(11.9,60){\circle*{1}}
1825: \put(128.1,40){\circle*{1}}
1826: \put(128.1,60){\circle*{1}}
1827: \put(18,30){\circle*{1}}
1828: \put(18,70){\circle*{1}}
1829: \put(122,30){\circle*{1}}
1830: \put(122,70){\circle*{1}}
1831: \put(30.3,20){\circle*{1}}
1832: \put(30.3,80){\circle*{1}}
1833: \put(109.7,20){\circle*{1}}
1834: \put(109.7,80){\circle*{1}}
1835: \spline(30.3,80)(40,86)(60,89)(70,90)(80,89)(100,86)(109.7,80)
1836: \spline(18,70)(40,84)(60,89)(70,90)(80,89)(100,84)(122,70)
1837: \spline(11.9,60)(40,84)(60,88)(70,90)(80,88)(100,84)(128.1,60)
1838: \spline(11.9,40)(40,83)(60,87)(70,90)(80,87)(100,83)(128.1,40)
1839: \spline(18,30)(40,80)(60,87)(70,90)(80,87)(100,80)(122,30)
1840: \spline(30.3,20)(40,78)(60,87)(70,90)(80,87)(100,78)(109.7,20)
1841: \put(55,35){\makebox(0,0){$p_i$}}
1842: \put(85,35){\makebox(0,0){$p_{\bar{i}}$}}
1843: %\put(70,3){\makebox(0,0){Fig. 4.14}}
1844: \end{picture}
1845: \end{center}
1846: \caption{The bimodule $X(p_i) \otimes_{M_\infty} X(p_{\bar{i}})$}
1847: \label{a3}
1848: \end{figure}
1849:
1850: Let $y_0$ be an arbitrary element of $X(p_i) \otimes_{M_\infty} X(p_j)$. This
1851: $y_0$ is considered as an element in the bimodule in Figure \ref{a3}. We
1852: denote it by $y_1$. Then, we attach $x \in {\rm Hom}(p_i \otimes p_j, p_k)$ to
1853: this $y_1$ by using two tubes $p_i$, $p_j$. See Figure \ref{a6}.
1854: \begin{figure}[htbp]
1855: \begin{center}
1856: \scalebox{0.45}{
1857: \includegraphics{a6.eps}}
1858: \end{center}
1859: \caption{}
1860: \label{a6}
1861: \end{figure}
1862: \begin{figure}[htbp]
1863: \begin{center}
1864: \scalebox{0.9}{
1865: \includegraphics{a7.eps}}
1866: \end{center}
1867: \caption{}
1868: \label{a7}
1869: \end{figure}
1870: The central part of Figure \ref{a6} can be viewed as in Figure \ref{a7}.
1871: Hence, we get a morphism in ${\rm Hom}(X(p_i) \otimes_{M_\infty} X(p_j),
1872: X(p_k))$. This induces a functor $G: \; {\cal D}(\Delta) \longrightarrow
1873: {\cal M}_\infty$. It is not difficult to see that $F$ and $G$ are
1874: functors, which preserve the operations in modular categories, and inverse to
1875: each other.
1876:
1877: Hence, the category of $M_\infty$-$M_\infty$ bimodules obtained from the
1878: asymptotic inclusion ${\cal M}_\infty$ is equivalent to the tube
1879: system ${\cal D}(\Delta)$ constructed from $N \subset M$ as modular
1880: categories.
1881:
1882:
1883: \newpage
1884: \begin{thebibliography}{100}
1885: \bibitem{Atiyah}
1886: M. F. Atiyah, {\it Topological quantum field theories},
1887: Publ. Math. I.H.E.S. 68 (1989) 175---186.
1888: \bibitem{EK} D. E. Evans, Y. Kawahigashi,
1889: {\it Quantum symmetries on operator algebras}, Oxford University Press, 1998.
1890: \bibitem{EK1} D. E. Evans, Y. Kawahigashi,
1891: {\it On Ocneanu's theory of asymptotic inclusions for subfactors, topological
1892: quantum field theories and quantum doubles}, Internat. J. Math. 6 (1995)
1893: 205--228.
1894: \bibitem{EK2} D. E. Evans, Y. Kawahigashi,
1895: {\it Orbifold subfactors from Hecke algebras II}, Commun. Math. Phys. 196
1896: (1998) 331--361.
1897: \bibitem{FSS}
1898: L. Feher, A. Stipsicz and J. Szenthe ed., {\it Topological field theories and geometry of loop spaces},
1899: World Scientific, 1992.
1900: \bibitem{Iz}
1901: M. Izumi, {\it Applications of fusion rules to classification of subfactors},
1902: Publ. RIMS, 27, (1991), 953--994.
1903: \bibitem{Izumi1}
1904: M. Izumi, {\it The structures of sectors associated with the
1905: Longo-Rehren inclusions I. General theory}, Commun. Math.Phys. 213 (2000)
1906: 127---179.
1907: \bibitem{Izumi2}
1908: M. Izumi, {\it The structures of sectors associated with the
1909: Longo-Rehren inclusions II. Examples}, Rev. in Math.Phys. 13 (2001)
1910: 603---674.
1911: \bibitem{J}
1912: V. F. R. Jones, {\it A polynomial invariant for knots via von Neumann
1913: algebras}, Bull. of the AMS, 12 (1985), 103--112.
1914: \bibitem{Kerler}
1915: T. Kerler, {\it Genealogy of non-perturbative quantum-invariants
1916: of 3-manifold: the surgical family}, in ``Geometry and Physics''
1917: (Aarhus 1995), 503--547. Lect. Notes in Pure and Appl. Math.,
1918: 184, Dekker, New York, 1997.
1919: \bibitem{Kohno}
1920: T. Kohno, {\it Topological invariants for $3$-manifolds using representations
1921: of mapping class groups II: Estimating tunnel number of knots},
1922: Contemp. Math. 175 (1994) 193--217.
1923: \bibitem{LR}
1924: R. Longo and J. E. Roberts, {\it A theory of dimension}, $K$-theory 11 (1997),
1925: 103--159.
1926: \bibitem{Mueger}
1927: M. M\"uger, {\it From subfactors to categories and topology II.
1928: The quantum double of subfactors and categories}, preprint, math.CT/0111205,
1929: 2001.
1930: \bibitem{Ocneanu}
1931: A. Ocneanu, {\it Chirality for operator algebras}, 39---63, in
1932: ``Subfactors", ed. by H. Araki, et al., World Scientific, 1994.
1933: \bibitem{O}
1934: A. Ocneanu, {\it Operator Algebras, Topology and Subgroups of Quantum
1935: Symmetry--Construction of Subgroups of Quantum Groups--}, 235---263,
1936: written by Satoshi Goto(Sophia University) and Nobuya Sato(Osaka Prefecture
1937: University), in ``Advanced Studies in Pure and Mathematics'' 31, Taniguchi
1938: Conference in Mathematics Nara '98, Mathematical Society of Japan, 2000.
1939: \bibitem{RT}
1940: N. Reshetikhin and V. G. Turaev, {\it Invariants of $3$-manifolds via link polynomials and quantum groups}, Invent. Math. 103 (1991), p. 547--597.
1941: \bibitem{Roberts}
1942: J. Roberts, {\it Skein theory and Turaev-Viro invariants}, Topology 34 (1995),
1943: 771--787.
1944: %\bibitem{Rolfsen}
1945: %D. Rolfsen, {\it Knots and links},
1946: %Publish or Perish, Berkeley 1976.
1947: \bibitem{SatoWakui}
1948: N. Sato and M. Wakui, {\it (2+1)-dimensional topological quantum field theory
1949: with Verlinde basis and Turaev-Viro-Ocneanu invariants of 3-manifolds},
1950: preprint 2000.
1951: \bibitem{SuzukiWakui}
1952: K. Suzuki and M. Wakui,
1953: {\it On the Turaev-Viro-Ocneanu invariant of $3$-manifolds derived from the
1954: $E_6$-subfactor}, Kyushu J. Math. 56 (2002) 59--81.
1955: \bibitem{Takata}
1956: T. Takata, {\it Invariants of $3$-manifolds associated with quantum groups and Verlinde's formula},
1957: Publ. RIMS Kyoto Univ. 28 (1992) 139--167.
1958: \bibitem{Turaev}
1959: V. G. Turaev, {\it Quantum invariants of knots and 3-manifolds}, Walter
1960: de Gruyter 1994.
1961: \bibitem{TuraevViro}
1962: V. G. Turaev and O. Ya Viro,
1963: {\it State sum invariants of 3-manifolds and quantum $6j$-symbols},
1964: Topology 31 (1992) 865---902.
1965: \bibitem{Verlinde}
1966: E. Verlinde, {\it Fusion rules and modular transformations in 2D conformal field theory}
1967: , Nucl. Phys. B300 (1988) 360--376.
1968: \bibitem{Wakui}
1969: M. Wakui, {\it Fusion algebras for orbifold models (a survey)},
1970: in \lq\lq Topology, geometry and field theory" edited by K. Fukaya, M. Furuta, T. Kohno
1971: and D. Kotschick, World Scientific, 1994, 225--235.
1972: \bibitem{Witten}
1973: E. Witten, {\it Quantum field theory and Jones polynomial},
1974: Commun. Math. Phys. 121 (1989) 351--399.
1975:
1976: \end{thebibliography}
1977:
1978:
1979: \end{document}
1980: