1: %Date 26 June 2006
2: \documentclass[reqno]{amsart} \usepackage{amscd}
3: \usepackage{epsfig}
4: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5: \newtheorem{theorem}{Theorem}[section]
6: \newtheorem{proposition}[theorem]{Proposition}
7: \newtheorem{lemma}[theorem]{Lemma}
8: \newtheorem{corollary}[theorem]{Corollary}
9: %
10: \newtheorem{alphthm}{Theorem}
11: \renewcommand{\thealphthm}{\Alph{alphthm}}
12: %
13: \theoremstyle{definition}
14: \newtheorem{definition}[theorem]{Definition}
15: \newtheorem{example}[theorem]{Example}
16: %
17: \theoremstyle{remark} \newtheorem{remark}[theorem]{Remark}
18: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
19: \numberwithin{equation}{section}
20: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
21: \newcommand{\be}{\begin{eqnarray}}
22: \newcommand{\ee}{\end{eqnarray}}
23: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
24: \newcommand{\field}[1]{\ensuremath{\mathbb{#1}}}
25: \newcommand{\CC}{\field{C}}
26: \newcommand{\DD}{\field{D}}
27: \newcommand{\HH}{\field{H}}
28: \newcommand{\PP}{\field{P}}
29: \newcommand{\RR}{\field{R}}
30: \newcommand{\TT}{\field{T}}
31: \newcommand{\ZZ}{\field{Z}}
32: \newcommand{\MM}{\field{M}}
33: \newcommand{\NN}{\field{N}}
34: %
35: \newcommand{\LL}{L^+} % positive light cone
36: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
37: %
38: % operator names
39: \renewcommand{\d}{\operatorname{d}}
40: \DeclareMathOperator{\id}{id}
41: \DeclareMathOperator{\Op}{\hat{\mathcal O}}
42: \DeclareMathOperator{\I}{I}
43: \DeclareMathOperator{\D}{D(SU(2))}
44: \DeclareMathOperator{\T}{{\mathcal T}}
45: \DeclareMathOperator{\Tr}{Tr}
46: \DeclareMathOperator{\Hom}{Hom}
47: \DeclareMathOperator{\Tor}{Tor}
48: \DeclareMathOperator{\Ext}{Ext}
49: \DeclareMathOperator{\Ker}{Ker}
50: \DeclareMathOperator{\im}{Im}
51: \DeclareMathOperator{\SL}{SL}
52: \DeclareMathOperator{\PSL}{PSL}
53: \DeclareMathOperator{\SO}{SO}
54: \DeclareMathOperator{\SU}{SU}
55: \DeclareMathOperator{\so}{\mathfrak so}
56: %
57: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
58: \begin{document}
59:
60: \title[Particles and GFT]{Quantum Gravity with Matter \\ via Group Field Theory}
61: %
62: \author{Kirill Krasnov} \address{School of Mathematical Sciences \\
63: University of Nottingham \\ Nottingham, NG7 2RD, UK}
64:
65: %
66: \begin{abstract} A generalization of the matrix model idea to quantum
67: gravity in three and higher dimensions is known as group field theory
68: (GFT). In this paper we study generalized GFT models that can be
69: used to describe 3D quantum gravity coupled to point particles. The
70: generalization considered is that of replacing the group leading to pure
71: quantum gravity by the twisted product of the group with its dual
72: --the so-called Drinfeld double of the group. The Drinfeld double is
73: a quantum group in that it is an algebra that is both non-commutative
74: and non-cocommutative, and special care is needed to define group
75: field theory for it. We show how this is done, and study the
76: resulting GFT models. Of special interest is a new topological model
77: that is the ``Ponzano-Regge'' model for the Drinfeld double. However, as we show,
78: this model does not describe point particles. Motivated by the
79: GFT considerations, we consider a more general class of models that
80: are defined using not GFT, but the so-called chain mail techniques.
81: A general model of this class does not produce 3-manifold invariants,
82: but has an interpretation in terms of point particle Feynman diagrams.
83: \end{abstract}
84: %
85: \maketitle
86: %
87: %
88: \section{Introduction}
89: \label{sec:intr}
90:
91: Recently a question of coupling of quantum gravity to matter has been revisited in works
92: \cite{Barrett,Freidel}. Both papers deal with quantum gravity in 2+1 dimensions, and analyze how
93: matter Feynman diagrams are modified when the quantum gravity effects are
94: taken into account. Both work contain results that are intriguing. However, certain
95: conceptual issues were left unclear. Thus, work \cite{Barrett} has suggested
96: that the effect of quantum gravity is in modifying the measure of integration
97: $\prod_v dx_v$ in a Feynman amplitude given by:
98: \be\label{coord-rep}
99: {\mathcal Z} = \int \prod_v dx_v \prod_{\langle vv'\rangle} G(x_v-x_{v'}).
100: \ee
101: Here one integrates over the positions of vertices $x_v$ of a Feynman graph, $G(x-y)$ is
102: the particle propagator in coordinate representation, and $\langle vv'\rangle$ denotes
103: edges of the Feynman diagram. In this scheme quantum gravity is only responsible for
104: a modification of the $\prod_v dx_v$ measure, and the particle content as well as the
105: Lagrangian must be specified independently. In particular, one is not forced to using any particular
106: propagator. Thus, this proposal is that of ``quantum gravity + matter''.
107: Work \cite{Freidel}, on the other hand, showed that particular quantum
108: gravity amplitudes can be interpreted as particle Feynman diagrams, provided
109: one makes a special choice of the particle propagator. Thus, in this
110: interpretation ``quantum gravity = matter''. These two interpretations
111: seem to be in conflict. This paper is devoted to an analysis of these and
112: related issues. We will argue that both interpretations are correct.
113:
114: The basic philosophy that is to be pursued in the present paper is as follows.
115: Consider the momentum representation Feynman amplitude of some
116: theory. The precise definition of the theory is unimportant for us at
117: this stage. We shall restrict our attention to general aspects common
118: to Feynman amplitudes of all theories. Thus, the amplitude is given by:
119: \be\label{mom-rep}
120: {\mathcal Z} = \int \prod_{\langle vv'\rangle} dp_{\langle vv'\rangle} G(p_{\langle vv'\rangle})
121: \prod_v \delta(\sum_{v'} p_{\langle vv'\rangle}).
122: \ee
123: Here the integrals are taken over the momenta $p_{\langle vv'\rangle}$ on each
124: edge of the digram; in addition, one has a momentum conservation law for each vertex.
125: We have to modify the Feynman amplitude \eqref{mom-rep} so that
126: it describes point particles. Point particles in 2+1 dimensions are
127: conical singularities in the metric. As we shall see, this has the effect that
128: particle's momentum is {\it group-valued}. For point particles in
129: Euclidean 3-dimensional space, which is what we consider in
130: this work, the relevant group is $\SU(2)$. The proposal to be developed in this paper is
131: that point particles and processes involving them can still be described by Feynman diagrams \eqref{mom-rep},
132: but the momentum $p$ on each edge of the diagram should be taken group-valued. The rest of
133: the paper is devoted to various tests of this proposal.
134:
135: Some point particle theories we consider are topological; we refer the
136: reader to the main body of the paper for the precise definition of
137: what this means. As we shall see, these topological particle theories
138: will be produced by the GFT. Other theories we consider are
139: non-topological; GFT is of no direct relevance here. However, the
140: techniques developed in the GFT context (such as the chain mail, see
141: below) will prove extremely useful for these theories as well. This
142: is how the GFT is omnipresent throughout the paper.
143:
144: We start by reminding the reader how point particles in 3D are described. For simplicity,
145: this paper deals only with the case of $3+0$-dimensional spacetime, that is, with 3D metrics
146: of Euclidean signature, and the cosmological constant is set to zero.
147:
148: \subsection{Point particles}
149:
150: In 3 spacetime dimensions presence of matter has much more drastic consequences than in 4D:
151: the asymptotic structure of spacetime gets modified. This has to do with the fact that the vacuum Einstein
152: equations are much stronger in 3D: they require that the curvature is constant (zero)
153: everywhere. Then an asymptotically flat spacetime is flat everywhere. When matter is
154: placed inside it introduces a conical defect that can be felt at infinity. The modification
155: due to matter is in the leading order, not in the next to leading as in higher dimensions.
156: This has the consequence that the total mass in spacetime is proportional to the deficit angle
157: created at infinity. Because the angle deficit cannot increase $2\pi$, the mass is
158: bounded from above. All this is very unlike the case of higher dimensions, and makes
159: the theory of matter look rather strange and unfamiliar. But such strange features are
160: at the end of the day responsible for a complete solubility of the theory.
161:
162: Point particles give the simplest and most natural form of matter to consider in 3D. A point
163: particle creates a conical singularity at the point where it is placed. The angle deficit
164: $2\theta$ at the tip of the cone is related to particle's mass as $4\pi M G=\theta$. Point
165: particles move along geodesics. Their worldliness are lines of conical singularities.
166:
167: To make this description more explicit, let us consider how a space containing a line
168: of conical singularity can be obtained. This is achieved by performing
169: a rotation of
170: space, and identifying points related by this rotation. A rotation is described by an
171: element $g(2\theta)\in\SU(2)$, which we choose to parameterize as follows:
172: \be
173: g(\theta)=e^{i \theta \vec{n}\cdot\vec{\sigma}}= \cos{\theta} + i \vec{n}\cdot\vec{\sigma} \sin{\theta}.
174: \ee
175: Here $\vec{n}: \vec{n}\cdot\vec{n}=1$ is a unit vector. It describes the direction of
176: the axis of rotation, and thus the direction of particle's motion. The angle $2\theta$
177: specifies the amount of rotation, and thus gives the mass of the particle. The reason for
178: choosing $\theta$ and not $2\theta$ for parameterizing the particle's mass will become clear below.
179: An equivalent description in more mathematical terms is to say that a particle of a given mass $\theta$
180: is described by a group element $g\in C_\theta$, where $C_\theta$ is a conjugacy class in $\SU(2)$.
181: Non-trivial conjugacy classes $C_\theta, \theta\not=0$ in $\SU(2)$ are spheres $S^2$
182: and a point in $C_\theta$ describes the direction of motion. Thus, the information about
183: both the mass and the direction of motion is encoded in $g$. This is why this quantity plays
184: the role of momentum of a point particle. Note that the group-valued nature
185: of particle's momentum automatically incorporates the bound on its mass. Note also
186: that not only the upper, but the lower bound is naturally imposed as well --
187: the $\SU(2)$ valued momenta are those of non-negative mass particles.
188:
189: Once points of $\RR^3$ related by a rotation $g$ are identified, we obtain a space with a
190: line of conical singularity in the direction $\vec{n}$. The angle deficit is $2\theta$.
191: The space has zero curvature everywhere except along particle's worldline. One can compute the
192: holonomy of the spin connection around the worldline and verify that it is equal to $g(\theta)$. It is for
193: this reason that the quantity $\theta$ equal to half the angle deficit gives a more convenient measure
194: parameterization of the mass. Let us now consider two point particles, with group elements $g_1,g_2$
195: describing them. The holonomy around the pair is the product of holonomies, and the mass
196: parameter $\theta$ of the pair is given by:
197: \be\label{mass-comb}
198: \cos{\theta} = \frac{1}{2} \Tr(g_1 g_2) = \cos{\theta_1}\cos{\theta_2} - \vec{n_1}\cdot\vec{n_2}
199: \sin{\theta_1} \sin{\theta_2}.
200: \ee
201: Note that it does matter now in which order the product of holonomies is taken, for
202: $g_1 g_2\not= g_2 g_1$ as we are dealing with elements of the group. But in determining the mass
203: of the combined system this ambiguity is irrelevant, for the trace is taken. Let us
204: look at \eqref{mass-comb} in more detail. In case $\vec{n_1}$ is parallel to $\vec{n_2}$, this formula
205: reduces to: $\theta=\theta_1+\theta_2$. When particles move in opposite directions we have:
206: $\theta=|\theta_1-\theta_2|$. Thus, the angle $\theta$ is more properly interpreted as
207: the magnitude of particle's momentum, not as its kinetic energy.
208:
209: It is instructive to compare \eqref{mass-comb} to the usual law of addition of vectors in $\RR^3$.
210: Thus, consider two particles with momenta $\vec{p_1}, \vec{p_2}$. Define: $m=|\vec{p}|, \vec{n}=\vec{p}/|\vec{p}|$.
211: Then we have:
212: \be\label{flat-comb}
213: m^2 = m_1^2 + m_2^2 + 2 \vec{n_1}\cdot\vec{n_2} m_1 m_2.
214: \ee
215: This formula for addition of momenta is to be compared with \eqref{mass-comb}. Indeed, as $\theta=4\pi MG$,
216: \eqref{mass-comb} can be expanded in powers of $G$. This gives exactly \eqref{flat-comb}, modified by
217: $O(G^2)$ corrections:
218: \be
219: M^2 = M_1^2 + M_2^2 + 2 \vec{n_1}\cdot\vec{n_2} M_1 M_2 + \\ \nonumber
220: \frac{4\pi^2 G^2}{3}\left( M^4 - M_1^4 - M_2^4 - 3M_1^2 M_2^2 - 4 \vec{n_1}\cdot\vec{n_2} M_1 M_2 (M_1^2+M_2^2)
221: \right) +O(G^4).
222: \ee
223: Thus, the fact that the law \eqref{mass-comb} of addition of momenta has the correct limit as $G\to 0$
224: gives additional support to our identification of $g$ with particle's momentum.
225:
226: Having understood why particle's momentum becomes group valued, let us remind the reader
227: some basic facts about the Ponzano-Regge model
228: \cite{PR} of 3D quantum gravity, and how point particles can be
229: coupled to it.
230:
231: \subsection{Ponzano-Regge model in presence of conical singularities}
232:
233: In Ponzano-Regge model,
234: one triangulates the 3-manifold in question (``spacetime''), and assigns
235: a certain amplitude to each triangulation. This amplitude turns out to be triangulation
236: independent and gives a topological invariant of the 3-manifold. Let us
237: remind the reader how the Ponzano-Regge amplitude \eqref{PR} can be
238: obtained from the gravity action. When written in the first order formalism, zero cosmological constant
239: 3D gravity becomes the so-called BF theory. The action is given by:
240: \be\label{bf}
241: S[{\bf e},{\bf a}]=\int_M {\rm Tr}({\bf e}\wedge {\bf f}).
242: \ee
243: Here $\bf e$ is a Lie algebra valued one form, and $\bf f$ is the curvature of a $G$-connection $\bf a$
244: on $M$. The path integral of this simple theory reduces to an integral over flat connections:
245: \be\label{Z-PR}
246: {\mathcal Z}[M]=\int {\mathcal D}{\bf e} {\mathcal D}{\bf a}\,\, e^{iS[{\bf e},{\bf a}]}=\int {\mathcal D}{\bf a}\,\,
247: \delta({\bf f}).
248: \ee
249: A discretized version of this last integral can be obtained as follows. Let us triangulate the
250: manifold $M$ in question. Let us associate with every edge of the dual triangulation the
251: holonomy matrix for the connection $\bf a$. This holonomy matrix is an element of the group $G$. The
252: product of these holonomies around the dual face must give the trivial element (because the connection is flat).
253: Thus, one can take the $\delta$-function of a product of group elements around each dual face,
254: multiply these $\delta$-functions and then integrate over the group elements on dual edges.
255: This is the discretized version of \eqref{Z-PR}. By decomposing the $\delta$-function on the
256: group $G=\SU(2)$ into characters, it is easy to show that this
257: procedure gives the following amplitude:
258: \be\label{PR-n}
259: \sum_{\{j_e\}} \prod_e {\rm dim}(j_e) \prod_t (6j).
260: \ee
261: It turns out, however, that this amplitude is not yet triangulation
262: independent. In order to make it such one has to introduce a certain
263: formal prefactor. Thus, let us consider the following sum over
264: irreducible representations of $\SU(2)$:
265: \be\label{eta}
266: \eta^{-2}:= \sum_j {\rm dim}_j^2.
267: \ee
268: Note that this sum diverges and the quantity $\eta$ defined
269: by it is formal. The sum above is equal to the $\delta$-function on the
270: group evaluated at the identity element. To make the amplitude
271: \eqref{PR-n} triangulation independent one has to consider the
272: following multiple of it:
273: \be
274: \label{PR}
275: \eta^{2V} \sum_{\{j_e\}} \prod_e {\rm dim}(j_e) \prod_t (6j),
276: \ee
277: where $V$ is the number of vertices in the triangulation used to
278: obtain \eqref{PR-n}. The amplitude \eqref{PR} is now (formally,
279: because of a prefactor that is actually equal to zero) triangulation
280: independent.
281:
282: The origin of this formal prefactor is now well-understood, see
283: \cite{Freidel-div}. It has to do with the fact that the action
284: \eqref{bf} in addition to the usual $\SU(2)$ gauge symmetry has a
285: non-compact gauge-symmetry. Indeed, because of
286: the Bianchi identity $d_A F=0$ the action is invariant under the
287: following transformation: $B\to B+ d_A \phi$. Integration over $B$
288: in the path integral includes integration over these gauge directions
289: and produces infinities that has to be cancelled by the zero prefactor
290: in \eqref{PR}. A more systematic procedure for dealing with these
291: divergences was developed in \cite{Freidel-div,Freidel-CS} and
292: requires a certain gauge fixing procedure. We shall not consider such
293: a gauge fixing here and will proceed at a formal level, just keeping
294: track of the (formal) pre-factors $\eta$. All amplitudes can be
295: rendered well-defined by passing to the quantum group context, that
296: is, by considering the quantum group $\SU_q(2)$ instead of
297: $\SU(2)$. When $q$ is a root of unity the cut-off on the spin present
298: in the representation theory of $\SU_q(2)$ makes the sum that defines
299: $\eta^{-2}$ finite, and all the invariants well-defined. One should
300: always have this passage to $\SU_q(2)$ in mind when considering the
301: ill-defined Ponzano-Regge model expressions.
302:
303: We have explained how the amplitude in \eqref{PR} can be obtained from an integral over flat connections.
304: It is now easy to add a point particle. Consider the case when a point particle is present and (a segment of)
305: its worldline
306: coincides with one of the edges of the triangulation of $M$. It will then create a conical singularity
307: along this edge. The product of holonomies around the face dual to this edge is not
308: trivial anymore. Instead it will lie in a conjugacy class determined by
309: the mass of the particle (by its deficit angle). Thus, for dual edges that encircle particles'
310: worldline one should put a $\delta$-function concentrated on the conjugacy class determined by
311: the particle. This gives a modified amplitude, with presence of the point particle taken
312: into account. Let us introduce a $\delta$-function that is picked on a particular conjugacy
313: class $\theta$:
314: \be
315: \delta_\theta(x) := \int dg \delta_e(g h_\theta g^{-1} x) = \sum_j \chi_j(h_\theta) \chi_j(x).
316: \ee
317: Here $\delta_e(x)$ is the usual $\delta$-function on the group picked
318: at the identity element. The sum in the second formula is taken over
319: all irreducible representations of the group $\SU(2)$, and
320: $\chi_j(g)$ are the characters. The special group element $h_\theta$ is a representative
321: of the conjugacy class $\theta: h_\theta={\rm diag}(\exp{i\theta},\exp{-i\theta})$. Note that when $\theta=0$,
322: the above $\delta$-function reduces to the usual $\delta$-function picked at the identity. Let us now
323: assume that a point particle is present along all the edges of the triangulation, and
324: denote the angle deficit on the edge $e$ by $\theta_e$. The modified quantum
325: gravity amplitude in the presence of particles is given by:
326: \be\label{PR-pp}
327: \sum_{\{j_e\}} \prod_e \chi_{j_e}(h_{\theta_e}) \prod_t (6j).
328: \ee
329: If particles are present along all the edges, there is no need to
330: introduce a formal $\eta$ prefactor as the amplitude is not required
331: to be triangulation independent. If the particle is only present
332: along some of the edges, one gets what can be called an observable of
333: Ponzano-Regge model, see \cite{Barrett-Obs}. In this case pre-factors
334: of $\eta$ for vertices away from the particle edges are necessary to
335: ensure a partial triangulation independence, see \cite{Barrett-Obs}
336: for more details.
337:
338: As it was explained in \cite{Barrett} and \cite{Freidel}, the
339: amplitude \eqref{PR-pp}
340: is the response of the quantized gravitational field to particle's presence. In the setting
341: described the particle is present as en external source in the gravitational path
342: integral. The natural question is if one can integrate over the
343: particle degrees of freedom and thus have a quantum theory of particles coupled to
344: quantum gravity. In other words, the question is how to ``second quantize'' the
345: particle degrees of freedom. Some sketches on how this can be done were presented in both
346: \cite{Barrett} and \cite{Freidel}. What is missing in these papers is a principle
347: that would generate Feynman diagrams for both gravity and particle degrees of
348: freedom. This paper is aimed at filling this gap. As we shall see, the
349: Feynman amplitudes for some of the theories (namely the topological
350: ones) can be generated by GFT. To understand
351: how this is possible, let us remind the reader some basic facts about the usual field
352: theory Feynman diagrams, and, in particular, about a certain duality transformation that can be
353: performed on a diagram. Let us remark that for the most part this paper deals with vacuum
354: (closed) Feynman diagrams only. A generalization
355: to open diagrams that are essential for doing the scattering theory will be
356: left for future work.
357:
358: \subsection{Momentum representation and duality}
359:
360: In addition to the coordinate \eqref{coord-rep} and momentum \eqref{mom-rep} representations
361: of Feynman diagrams there is another, the so-called dual representation. Let us
362: remind the reader how the dual representation is obtained in the case of
363: two spacetime dimensions. To get the dual formulation we need to endow
364: Feynman graphs with an additional ``fat'' structure. Namely, each edge
365: of the digram must be replaced by two lines. The fat structure is equivalent
366: to an ordering of edges at each vertex. The lines of the fat graph are then
367: connected at vertices as specified by ordering. This structure is also
368: equivalent to specifying a set of 2-cells, or faces of the diagram. Once this
369: additional structure is introduced, one can define the {\it genus} $G$ of the diagram
370: to be given by the Euler formula: $2-2G=V-E+F$, where $V,E,F$ are the numbers
371: of vertices, edges and faces of the diagram correspondingly. Note that the same
372: Feynman diagram can be given different fat structures and thus have different
373: genus. For example, the diagram with 2 tri-valent vertices and 3 edges (the $\theta$-graph)
374: can correspond to both genus zero and genus one after the fat structure is
375: specified.
376:
377: Having introduced the fat structure, we can solve all the momentum
378: conservation constraints. It is simplest to do it in the case of $G=0$, so let us
379: specialize to this situation. Let us introduce the new momentum variables,
380: one for every face of the diagram. Let us denote these by $p_f$. Given a pair
381: $\langle ff'\rangle$ of adjacent faces, there is a unique edge $\langle vv'\rangle$
382: that is a part of the boundary of both of them. A relation between the original
383: momentum variables $p_{\langle vv'\rangle}$ and the new variables $p_f$ is then given by:
384: \be\label{duality}
385: p_{\langle vv'\rangle}=p_f - p_{f'}.
386: \ee
387: Let us check that the number of new variables is the number of old variables minus the
388: number of constrains. The number of new variable is equal to $F-1$: the number of faces minus one.
389: We have to subtract one because everything depends on the differences only and does
390: not change if we shift all the variables by the same amount. We have $F-1=E-(V-1)-2G=E-(V-1)$,
391: where we have used our assumption that $G=0$. Thus, the number of new variables equals
392: to the number of the original variables, minus the number of $\delta$-functions,
393: equal to $V-1$ because one $\delta$-function is always redundant. In case $G\not=0$
394: one has to introduce an extra variable for each independent non-contractible loop
395: on the surface, whose number is $2G$. Thus, having expressed the momentum variables
396: via $p_f$ using \eqref{duality}, we have automatically solved
397: all the conservation constraints. The Feynman amplitude becomes:
398: \be\label{dual-rep}
399: \int \prod_f dp_f \prod_{\langle ff'\rangle} G(p_f - p_{f'}).
400: \ee
401:
402: We shall refer to \eqref{dual-rep} as the dual representation of a 2D Feynman
403: diagram. One can perform a similar duality transformation in higher dimensions.
404: The case relevant for us here is that of 3D, so let us briefly analyze what happens.
405: Let us consider a vacuum (closed) Feynman diagram and introduce an
406: additional structure that specifies faces, as well as an ordering of these faces around edges.
407: For example, let us consider Feynman graphs
408: whose edges and vertices are those of a {\it triangulation} of spacetime. As we
409: shall soon see, such Feynman diagrams are natural if one wants to interpret the
410: state sum models of 3D quantum gravity in particle terms. The faces of such
411: ``triangulation'' diagrams are then the usual triangles. As in 2D, let us assign
412: a variable $p_f$ to each face. The original edge momentum variable is then expressed
413: as an appropriate sum of the face variables, for all the faces that share the
414: given edge:
415: \be
416: p_e = \sum_{f: e\in f} p_f \epsilon_f.
417: \ee
418: One sums the face variables $p_f$ weighted with sign $\epsilon_f$;
419: a precise convention is unimportant for us at the moment. The number of the
420: original variables is $E-(V-1)$, and due to the fact that the Euler characteristic
421: of any 3-manifold is equal to zero, we get: $E-(V-1)=F-(T-1)$, which means that
422: one in addition has to impose $T-1$ constraints for the new variables.
423: Thus, in the original momentum representation we had $E$ edge variables
424: together with the momentum conservation constraints for each vertex.
425: In the dual formulation one has a momentum variable for each {\it dual edge},
426: and one conservation constraint for each dual vertex. We had to switch to the
427: dual lattice in order to solve the $\delta$-function constraints.
428:
429:
430: \subsection{Point particle Feynman diagrams}
431:
432: Here we further develop our proposal of
433: modifying Feynman diagrams by making the momentum group valued. Thus, let
434: us consider some theory of point particles in 3D that generates Feynman
435: diagrams. A form of the action is unimportant for us. Our starting point
436: will be diagrams in the momentum representation. As we have already described, the most
437: unusual feature of point particles in 3D is that their momentum is group-valued and thus non-commutative.
438: Thus, point particles in 3D behave unlike the standard relativistic fields in Minkowski spacetime.
439: In spite of this, one can still consider Feynman diagrams. Indeed, in the momentum
440: formulation \eqref{mom-rep} one only needs to specify what the propagator is, what the integration
441: measure $dp$ is, and what replaces the momentum conservation $\delta$-functions. All
442: this objects exist naturally for the 3D point particles. The propagator $G(p)$ is
443: some function on the group $\SU(2)$, which we shall leave unspecified for now. The
444: integration measure is the Haar measure on the group. The momentum conservation is
445: more subtle. The conservation constraint becomes the
446: condition that the product of the group elements $g_{\langle vv'\rangle}$ for all $v'$
447: is equal to the identity group element. However, it now does matter in which
448: order the group elements are multiplied. Thus, some additional structure is
449: necessary to define the momentum conservation constraints and Feynman amplitude as
450: a whole. We shall specify this additional structure below. Ignoring this issue for the moment we get:
451: \be\label{pp-ampl}
452: {\mathcal Z} = \int \prod_{\langle vv'\rangle} dg_{\langle vv'\rangle} G(g_{\langle vv'\rangle})
453: \prod_v \delta\left(\prod_{v'} g_{\langle vv'\rangle} \right).
454: \ee
455: Here $g_e=g_{\langle vv'\rangle}\in\SU(2)$ is the group element that describes particle's momentum on edge $e$,
456: $G(g)$ is a propagator which is a function on the group $\SU(2)$, and
457: $\delta(g)$ is the usual $\delta$-function on the group picked at the identity element.
458:
459: Let us now deal with the issue of defining the momentum conservation constraints. It is
460: clear that some additional structure is necessary. We have already seen an extra structure
461: being added to a Feynman diagram when we considered the dual formulation. In that case
462: an extra structure was introduced to solve the momentum constraints. It is clear that
463: exactly the same structure can be used to define these constraints in the case of
464: group-valued momenta. Thus, let us introduce an extra structure of faces, as well
465: as ordering of faces around each edge. This is equivalent to introducing the
466: structure of a dual complex. For simplicity we shall assume that the Feynman
467: diagram we started from is a triangulation of some manifold. The dual complex is then
468: the dual triangulation. Let us introduce new momentum variables: one for each face of the triangulation. The
469: original edge momentum is expressed as a product of the face variables:
470: \be\label{new-var}
471: g_e = \prod_{f: e\in f} g_f.
472: \ee
473: The order of the product here is important and is given by the ordering of the faces around the edge:
474: one multiples the holonomies across faces
475: to get the holonomy around the edge. The new face momentum variables $g_f$ solve
476: all the vertex momentum conservation constraints.
477: The amplitude \eqref{pp-ampl} in the dual formulation becomes:
478: \be\label{pp-dual}
479: {\mathcal Z} = \int \prod_f dg_f \prod_e G\left(\prod_{f: e\in f} g_f\right) \prod_t TC,
480: \ee
481: where $TC$ are the tetrahedral constraints that were shown to be necessary by counting the
482: variables in the previous subsection.
483:
484: \subsection{Tetrahedron constraints}
485:
486: It turns out that the tetrahedron constraints have a simple
487: meaning.\footnote{We thank Laurent Freidel for helping us with this
488: interpretation.} The constraints imply that the face variables are
489: not all independent. There is a gauge symmetry that acts at each
490: tetrahedron. At this stage it is convenient to pass to the dual
491: triangulation. Recall that each face is dual to an edge $e$ of the dual
492: triangulation, and each tetrahedron is dual to a vertex $v$. To write
493: down the action of the gauge transformation acting at $v$, let us orient all the
494: edges $e$ incident at this vertex to point away from $v$. Then the
495: action on $g_e$ is as follows: $g_e\to g_v g_e$, where $g_v$ is the
496: same for all edges $e$ incident at $v$. Fixing this gauge symmetry is
497: the same as imposing the tetrahedron constraints in
498: \eqref{pp-dual}. Below we shall see how these constraints can be dealt with.
499:
500: There is another possible interpretation that can be given to the constraints.
501: Let us consider a geometric tetrahedron in $\RR^3$.
502: The group valued face variables $g_f$ that we have introduced have the meaning of
503: an $\SU(2)$ transformation that has to be carried out when one goes from one tetrahedron
504: of the triangulation to the other. This group element can be represented as a product of
505: two group elements, one for each tet: $g_f = (g_f^t)^{-1} g_f^{t'}$, where $t,t'$ are the two
506: tetrahedra that share the face $f$. Consider now: $(g^t_f)^{-1} g^t_{f'}$. This group
507: element describes the rotation of one face $f$ into another $f'$, and thus
508: carries information about the dihedral angle between the two faces. Thus, the group variables
509: $g_f$ introduced to solve the momentum constraints carry information about the dihedral
510: angles between the faces of the triangulation. In particular, the total angle $2\pi-\theta_e$ on an edge
511: $e$ is just the sum of all the dihedral angles around this edge:
512: \be
513: 2\pi-\theta_e = \sum_{\langle ff'\rangle} \theta_{ff'},
514: \ee
515: where the sum is taken over the pairs $\langle ff'\rangle$ of faces that share the edge $e$.
516: This relation should be thought of as contained in the relation \eqref{new-var}.
517: Having understood the (partial) geometrical meaning of the face variables as encoding the dihedral
518: angles between the faces, we can state the (partial) meaning of the tetrahedron constraints. Recall
519: that 6 dihedral angles of a tetrahedron satisfy one relation. This relation is obtained as
520: follows. Consider the unit normals $\vec{n_i}, i=1,\ldots,4$ to all 4 faces of the tetrahedron.
521: Let us form a matrix $A_{ij}=\vec{n_i}\cdot\vec{n_j}$ of products of the normals. It is clear
522: that $\vec{n_i}\cdot\vec{n_j} = -\cos(\theta_{ij})$, and so the entries of $A_{ij}$ are
523: just the cosines of the dihedral angles. However, as the vectors $\vec{n_i}$ are 4 vectors in
524: $\RR^3$, they must be linearly dependent:
525: \be
526: \sum_i A_i \vec{n_i} = 0.
527: \ee
528: The coefficients $A_i$ here are just the areas of the corresponding faces. Because the normals
529: $\vec{n_i}$ are linearly dependent the determinant of the matrix $A_{ij}$ is equal to zero,
530: which is the sought constraint among the 6 dihedral angles
531: $\theta_{ij}$. This relation between the dihedral angles should be
532: thought of as contained in the tetrahedron constraints in \eqref{pp-dual}.
533:
534: Let us now consider some example theories to which the above strategy
535: can be applied.
536:
537: \subsection{Pure gravity from point particle Feynman diagrams}
538:
539: One special important case to be considered arises when particle's propagator is equal to the
540: $\delta$-function of the momentum. Had we been dealing with a usual theory in $\RR^3$ in
541: which the position and momentum representations are related by the Fourier transform
542: such a choice of the propagator would mean having $G(x-y)=1$ in the position space.
543: Thus, this theory is a trivial one as far as particles are concerned - there are no
544: particles. However, as we shall see in a moment, in the point particle context it
545: leads to a non-trivial and interesting amplitude.
546:
547: From the perspective of the Feynman amplitude \eqref{pp-ampl} it is clear that this case is a bit
548: singular. Indeed, the edge $\delta$-functions guarantee that all the holonomies
549: around edges are trivial. But then the vertex momentum conservation $\delta$-function
550: is redundant, and makes the amplitude divergent. To deal with this divergence, let
551: us consider a modified amplitude with no momentum conservation $\delta$-functions present.
552: An equivalent way of dealing with this problem is to introduce the
553: (formal) quantity $\eta$ given by \eqref{eta}.
554: Now we have to divide our amplitude by $\eta^{-2V}$, where $V$
555: is the number of vertices in the diagram. We can now consider the dual formulation
556: in which all the constraints are solved. Let us keep the factors of $\eta$, for they
557: are going to be an important part of the answer that we are about to get. The amplitude
558: in the dual formulation is given by:
559: \be
560: {\mathcal Z} = \eta^{2V} \int \prod_f dg_f \prod_e \delta\left(\prod_{f: e\in f} g_f\right) \prod_t TC.
561: \ee
562: Now it is easy to verify that the expression under the integral is
563: invariant under the gauge transformations that act at the vertices of
564: the dual triangulation. Thus, the tetrahedron constraints are
565: redundant and can be dropped. This does not introduce any
566: divergences, as the extra integrations performed if one drops the
567: constraints gives the volume of the gauge group to the power of the
568: number of tetrahedra. We always assume that the Haar measure on the
569: group is normalized so that the volume of the group is one. Thus, when
570: a gauge invariant quantity is integrated, the tetrahedron constraints
571: can be dropped.
572:
573: Thus, let us consider:
574: \be
575: {\mathcal Z} = \eta^{2V} \int \prod_f dg_f \prod_e \delta\left(\prod_{f: e\in f} g_f\right).
576: \ee
577: It is now possible to take all the group integrations explicitly. Indeed,
578: decomposing all the $\delta$-function into characters, and performing the group integrations,
579: one gets exactly the Ponzano-Regge amplitude \eqref{PR}, with the important prefactor of
580: $\eta^{2V}$ that is missing in \eqref{PR-n}. It is only when this prefactor is present that
581: the Ponzano-Regge amplitude is triangulation independent.
582:
583: Thus, we learn that the Ponzano-Regge amplitude \eqref{PR}, which, according to our previous
584: discussion, is equal to the quantum gravity partition function, can also be interpreted
585: as an amplitude for a point particle Feynman diagram with the particle's propagator given by
586: the $\delta$-function of the momentum. A few comments are in order. First, the Ponzano-Regge
587: amplitude is independent of the triangulation chosen. Therefore, if we now take the particle interpretation, the
588: Feynman amplitude is independent of a Feynman diagram. This could have been
589: expected, because particle's propagator given by the $\delta$-function essentially says that
590: no particle is present at all. This is why any Feynman diagram gives the same result.
591: The second comment is on why such two different
592: and seemingly contradictory interpretations are possible. Indeed, in one of the
593: pictures, we are dealing with the pure gravity partition function, and no particles
594: is ever mentioned. In the second picture, we consider point particles, whose momentum
595: became non-commutative group-valued. This happened due to the back-reaction of the
596: particle on the geometry. Point particle Feynman diagrams as integrals over
597: the group manifold take this back-reaction into account. Because pure gravity
598: does not have its own degrees of freedom, the only variables we have to integrate
599: over are those of the particles. Thus, particle's Feynman diagrams do give particle's
600: quantum amplitudes with quantum gravity effects taken into account. This is why no
601: separate path integration over gravity variables is necessary.
602:
603: The picture emerging is very appealing. Indeed, one has two different interpretations
604: possible. In one of them one considers pure gravity with its ``topological'' degrees
605: of freedom. The other picture gives a ``materialistic'' interpretation of the pure
606: gravity result in terms of point particles.
607:
608: It is now clear that a generalization is
609: possible. Indeed, one does not have to take the propagator to be the $\delta$-function.
610: For example, one can consider the $\delta$-function $\delta_\theta$ picked on some non-trivial
611: conjugacy class $\theta$. The conjugacy classes can be chosen to be independent for
612: different edges. Then in the dual formulation, one arrives at the Ponzano-Regge
613: amplitude \eqref{PR-pp} modified by the presence of particles.
614:
615: The next level of generalization would be to allow all possible values of $\theta$ on
616: every edge, and integrate over each $\theta_e$. This would mean integrating
617: over possible masses of point particles on every edge. Unlike the theory with a given
618: fixed set of angle deficits, which gives quantum gravity amplitudes modified by the presence
619: of a fixed configuration of point particles, theory that integrates over particle's
620: masses describes quantized particles together with quantized gravity. It is
621: our main goal in this paper to study what types of theories of this type are
622: possible. Thus, having understood that quantum gravity is described by
623: taking into account a modification of particle's momentum that is due to the
624: back-reaction, the question to ask is what fixes the theory that describes
625: the particles. In other words, what is a principle that selects particle's
626: propagator? What fixes valency of the vertices, that is the type of
627: Feynman diagrams that are allowed? In the usual field theory we have the
628: action principle that provides us with an answer to all these questions. In our
629: case there is no such principle known. The models that we have considered so far
630: are only known in the momentum representation or the corresponding dual.
631: In fact, because the momentum space is non-commutative
632: group manifold, one should expect the same to be true about the coordinate space.
633: Thus, one can expect that the original coordinate representation of theories
634: of the type we consider is in some non-commutative space. It has been argued recently that
635: even in the no-gravity limit $G\to 0$ this non-commutativity survives and leads to doubly special relativity, see
636: \cite{Freidel} as well as a more recent paper \cite{Majid-F} and references therein for more details.
637:
638: Thus, one possible direction would be to understand what is the coordinate
639: representation for point particle theories, and then postulate some
640: action principle that would fix the propagator and interaction type. However, in this paper
641: we shall instead consider the
642: dual momentum formulation as fundamental. The question we address is which
643: particle theories lead to ``natural'' dual representations. Indeed, to
644: arrive to the quantum gravity interpretation in terms of the Ponzano-Regge
645: amplitude we had to switch to the dual representation. It is the
646: dual representation that introduces a triangulation, and thus gives
647: some relation to geometry. Another argument in favor of the dual representation is
648: that Feynman diagrams in the original momentum representation are insensitive to
649: the topology of the manifold they are on. Thus, they do not contain
650: all of the degrees of freedom of the gravitational field, only the
651: local point particle ones. To describe the global degrees of freedom
652: one needs to have the extra structure that is available in the dual
653: formulation. Finally, even to define what the Feynman amplitude is in the case of a group-valued
654: momentum requires the dual representation. Indeed, to define the momentum conservation
655: constraints one needs the structure of faces of the diagram, which is only
656: available with the dual graph. All this makes it rather clear that the dual
657: formulation is indispensable for the description of point particles in 3 dimensions.
658: The logic of this paper will be to start from the formulation in terms of group field theory
659: and the dual graph and then give the Feynman diagram interpretation.
660:
661: With this idea in mind, we have to discuss how to obtain theories for
662: which a dual formulation is available. A natural way to do this is
663: using a generalization of the matrix model idea. Such a generalization
664: is available already for the Ponzano-Regge model, and is known
665: under the name of Boulatov theory \cite{Boulatov}. We shall describe
666: similar theories that incorporate point particles. Before we
667: do this, let us remind the reader some basic facts about the usual
668: matrix models.
669:
670: \subsection{Matrix models}
671:
672: Much of the activity in theoretical physics at the end of 1980's beginning of 1990's was
673: concentrated in the area of matrix models. The idea is that the perturbative
674: expansion of a simple matrix model given by the action:
675: \be
676: S[M]= N {\rm Tr}\left( \frac{1}{2} M^2 + V(M) \right),
677: \ee
678: where $M$ is, say, a hermitian $N\times N$ matrix, and $V(M)$ is some (polynomial) potential,
679: can be interpreted as a sum over discretized random surfaces. Such a sum, in turn, gives
680: the path integral of Euclidean 2D gravity. The above matrix model can be solved
681: by a variety of techniques and the predicted critical exponents match those obtained
682: by the continuous methods. The simple one matrix model
683: given above corresponds to the theory of pure 2D gravity, or, if one chooses the potential
684: $V(M)$ appropriately, gives 2D gravity coupled to the so called $(2,2m-1)$ matter. Couplings to
685: other types of $c<1$ matter are possible by considering the multi-matrix models, see e.g.
686: \cite{MM} and references therein for more details.
687:
688: More generally, one can consider a matrix model for an infinite set of matrices parameterized
689: by a continuous parameter that is customarily referred to as time $t$.
690: The corresponding matrix quantum mechanics
691: gives 2D gravity coupled to a single scalar field, i.e. the so-called $c=1$ matter,
692: and is given by the following action:
693: \be\label{mqm}
694: S[M(t)]=\int dt N {\rm Tr}\left( \frac{1}{2} \dot{M}^2+ \frac{m^2}{2} M^2 + V(M) \right).
695: \ee
696: Here the matrices $M(t)$ are functions of time, and $m$ is a ``mass'' parameter.
697: The matrix quantum mechanics \eqref{mqm} can be solved exactly. The results can be compared with
698: those obtained in the continuous approaches, with full agreement, see \cite{MM} for more details.
699:
700: \subsection{Field theory over a group manifold}
701:
702: The idea of generating random discretized manifolds from a matrix model was generalized to 3 dimensions in
703: \cite{Boulatov}. Here one considers a field $\phi(g_1,g_2,g_3)$
704: on three copies of a group manifold $G$. The field $\phi(g_1,g_2,g_3)$ should in addition
705: satisfy an invariance property: $\phi(g_1g,g_2g,g_3g)=\phi(g_1,g_2,g_3)$.
706: One should think of each argument of the field as
707: a generalization of the matrix index in 2D. One considers the following simple action:
708: \be\label{GFT-1}
709: S[\phi] = \frac{1}{2} \int dg_1 dg_2 dg_3 \,\,
710: \phi^2(g_1,g_2,g_3) + \frac{g}{4!} \prod_{i<j} \int dg_{ij} \prod_i \phi(g_{ij},g_{ik},g_{il}).
711: \ee
712: Here the first product in the interaction term is over pairs of indices $i,j=1,\ldots,4$, and
713: indices $i,j,k,l$ in the arguments of $\phi$ are such that $i\not=j\not=k\not=l$.
714: Using the Fourier analysis on the group one can expand function $\phi$ into the group matrix elements.
715: One then obtains the Feynman rules from the action written in the ``momentum'' representation.
716: The resulting sum is over triangulated 3-manifolds, or, more
717: precisely, pseudo-manifolds. Edges of the Feynman graph correspond to faces
718: of the triangulation, and Feynman graph vertices
719: corresponding to the tetrahedra. Edges of the complex get labeled by
720: irreducible $G$-representations $l_e$ and one has to sum over these. The perturbative expansion
721: is given by:
722: \be
723: \sum_{complexes} g^T \sum_{\{l_e\}} \prod_e {\rm dim}(l_e) \prod_t (6j).
724: \ee
725: The first product here is over the edges of the triangulation, and the second is over its 3-cells
726: (tetrahedra). The quantity $T$ is the total number of tetrahedra in a complex.
727: The quantity $(6j)$ is the Racah coefficient that depends on 6 representations.
728: The amplitude of each complex here is (almost, because of the missing
729: $\eta$ factors) the Ponzano-Regge amplitude \cite{PR}.
730: We shall give details of the calculation that leads to this result in the next section.
731:
732: Thus, the generalized matrix model of Boulatov \cite{Boulatov} gives a sum over random triangulated 3-manifolds
733: weighted by the exponent of the discrete version of the gravity action. Unlike the case of 2D matrix models
734: it was not possible to solve the Boulatov theory, and not so much has been learned from it about
735: 3D quantum gravity, see, however, \cite{Sum} for an interesting proposal for summing over the
736: 3-topologies. Similar theories have been considered in higher number of
737: dimensions as well. This more general class of models defined on the group manifold
738: and leading to sums over triangulated manifolds have received the name of group field
739: theories (GFT).
740:
741: \subsection{Outline and organization of the paper}
742:
743: We have seen that the
744: Ponzano-Regge amplitude has a point particle interpretation: it is the amplitude
745: for a theory of point particles with a propagator being the
746: $\delta$-function in the momentum space, and the amplitude is computed
747: in the dual momentum representation. In this paper we shall address the following question:
748: what kind of group field theory can give dual momentum representation for point particle theory
749: with a non-trivial propagator? As we shall see, the requirement that Feynman diagrams
750: generated by group field theory also admit an interpretation as dual of some other
751: Feynman diagrams severely restricts the possible choices of theories. There is essentially
752: a few possible choices, which we shall discuss in due course.
753:
754: The main idea of this paper is to consider a group field theory on the {\it Drinfeld double}
755: $\D$ of the group $\SU(2)$. To the best of our knowledge, the fact that point particles in 3D are
756: naturally described by $\D$ was discovered and explored in \cite{BM}. More recently, it
757: was used in \cite{Freidel-CS} to give the CS formulation of the Ponzano-Regge
758: model. As we shall see, a group field theory construction on $\D$ similar
759: to that of Boulatov \cite{Boulatov} can be performed. In some sense, the
760: idea of going from $\SU(2)$ group field theory to $\D$ one is analogous to
761: the way one introduces the matter degrees of freedom in 2D, where multi-matrix instead
762: of one-matrix models are considered. However, the analogy here is not
763: direct, for, as we shall find, the most obvious generalization does
764: not lead to a theory with point particles in it. Nevertheless, group
765: field theory considerations will allow us to develop certain methods
766: that will later be used to describe a general point particle theory.
767:
768: The organization of this paper is as follows. In the next section we
769: describe how Ponzano-Regge model is obtained from Boulatov
770: theory. Here we also give an abstract algebraic formulation of
771: Boulatov model that will be later used to define GFT for the Drinfeld
772: double. We describe the Drinfeld double of a classical group in
773: section \ref{sec:DD}. Here we discuss irreducible representations of
774: $\D$, characters, as well as an important projector property satisfied
775: by them. Group field theory for the Drinfeld double is defined in
776: section \ref{sec:gft}, and some of its properties are proved. Section
777: \ref{pp-models} defines more general particle models that are not
778: related to GFT. Interpretation of all the models is developed
779: in section \ref{sec:pp}. We conclude
780: with a discussion of the results obtained.
781:
782: Let us note that, apart from the works already mentioned, the question of coupling of quantum
783: gravity to matter was considered in paper \cite{Mikovic}. However, the approach taken in the
784: present paper is rather different.
785:
786: \section{Boulatov field theory over the group}
787: \label{sec:Boul}
788:
789: Boulatov theory \cite{Boulatov} can be formulated in several equivalent ways. One of
790: this formulation will be used for generalization to the Drinfeld double. Let us
791: first give the original formulation used in \cite{Boulatov}
792:
793: \subsection{Original Boulatov formulation}
794:
795: Let us consider a scalar field on the group manifold $\phi(g)$. For definiteness we take the group to
796: be $G=\SU(2)$ that corresponds to 3D Euclidean gravity.
797: However, one can take any (compact) Lie group. The resulting theory would still be topological, and be
798: related to the so-called BF theory for the group $G$. The
799: Fourier decomposition of the function $\phi$ is given by:
800: \be\label{b1}
801: \phi(g) = \sum_{l} \sum_{mn} t^l_{mn}(g) \phi^l_{mn}.
802: \ee
803: Here $t^l_{mn}(g)$ are the matrix elements in $l$'th representation, and $\phi^l_{mn}$ are the
804: Fourier coefficients. It is extremely helpful to introduce a graphic notation for this formula.
805: Denoting the Fourier coefficients by a box, and the matrix elements by a circle, both with two lines
806: for indices $m,n$ sticking out, we get:
807: \be
808: \phi(g) = \sum_l \,\,\, \lower0.23in\hbox{\epsfig{figure=four-1.eps, height=0.45in}}
809: \ee
810: The sum over $m,n$ is implied here. The graphical notation is much
811: easier to read than \eqref{b1}, and we shall write many formulas using
812: it in what follows.
813:
814: The field of Boulatov theory is on 3 copies of the group manifold. Analogous Fourier expansion is
815: given by:
816: \be
817: \phi(g_1,g_2,g_3) = \sum_{l_1,l_2,l_3} \,\,\, \lower0.25in\hbox{\epsfig{figure=four-3.eps, height=0.5in}}
818: \ee
819: The field is required to be symmetric:
820: \be
821: \phi(g_{\sigma(1)},g_{\sigma(2)},g_{\sigma(3)})=(-1)^{|\sigma|} \phi(g_1,g_2,g_3).
822: \ee
823: Here $\sigma$ is a permutation, and $|\sigma|$ its signature.
824: In addition, it is required to satisfy the following invariance property:
825: \be\label{sym}
826: \phi(g_1 g,g_2 g,g_3 g)=\phi(g_1,g_2,g_3).
827: \ee
828: Let us find consequences of this for the Fourier decomposition. Let us integrate the right hand side of
829: \eqref{sym} over the group. We are using the normalized Haar measure, so we will get $\phi(g_1,g_2,g_3)$
830: on the right hand side. On the left hand side we can use the formula \eqref{A-3} for the integral of three
831: matrix elements. Thus, we get:
832: \be\label{phi-1}
833: \phi(g_1,g_2,g_3) = \sum_{l_1,l_2,l_3} \,\,\,\lower0.5in\hbox{\epsfig{figure=phi-3-1.eps, height=1in}}
834: \ee
835: Let us now introduce a new set of Fourier coefficients $\tilde{\phi}^{l_1 l_2 l_3}_{m_1 m_2 m_3}$.
836: Graphically, they are defined by:
837: \be\label{tilde-phi}
838: \lower0.25in\hbox{\epsfig{figure=tilde-phi.eps, height=0.5in}}
839: \ee
840: We shall only use the new, modified set of coefficients
841: \eqref{tilde-phi} till the end of this section. Thus, we
842: shall omit the tilde. Our final expression for the Fourier decomposition is given by:
843: \be\label{phi-2}
844: \phi(g_1,g_2,g_3) = \sum_{l_1,l_2,l_3} \,\,\,\lower0.45in\hbox{\epsfig{figure=phi-3.eps, height=0.9in}}
845: \ee
846:
847: It is now straightforward to write an expression for the action in terms of the Fourier coefficients.
848: The action \eqref{GFT-1} is designed in such a way that there is always two matrix elements containing
849: the same argument. Thus, a repeated usage of the formula \eqref{A-2} gives:
850: \be\label{action-f}
851: S[\phi] = \frac{1}{2} \sum_{l_1 l_2 l_3}\!{}' \,\,\,\lower0.25in\hbox{\epsfig{figure=kinetic.eps, height=0.5in}}
852: + \\ \nonumber
853: \frac{g}{4!} \sum_{l_{ij}}\,\,\,\left( \begin{array}{ccc} l_{12} & l_{13} & l_{14} \\
854: l_{34} & l_{24} & l_{23} \end{array}\right) \,\,\, \lower0.5in\hbox{\epsfig{figure=inter.eps, height=1in}}
855: \ee
856: Here the first sum is taken only over sets $l_1,l_2,l_3$ that satisfy
857: the triangular inequalities, and this is reflected in a prime next to the sum symbol.
858: To arrive to this expression we have used the normalization of the intertwiner given by \eqref{A-theta}.
859: We have also introduced the so-called Racah coefficient, or the $6j$-symbol that is denoted by
860: two rows of spins in brackets.
861:
862: It is now easy to derive the Feynman rules. The propagator and the vertex are given by, correspondingly:
863: \be
864: \lower0.23in\hbox{\epsfig{figure=prop.eps, height=0.45in}}
865: \ee
866: where the black box denotes the sum over permutations, and
867: \be
868: g \left( \begin{array}{ccc} l_{12} & l_{13} & l_{14} \\
869: l_{34} & l_{24} & l_{23} \end{array}\right)
870: \,\,\, \lower0.45in\hbox{\epsfig{figure=vertex.eps, height=0.9in}}
871: \ee
872: Note now that each Feynman graph of this theory is a dual skeleton of some simplicial 3D manifold. Indeed,
873: vertices of the graph correspond to simplices, edges corresponds to faces, and faces
874: (closed loops) corresponds to edges. A proper way to
875: describe the 3-manifolds arising is in terms of pseudo-manifolds. It will be presented below.
876: Each Feynman amplitude is weighted by {\it almost} the Ponzano-Regge amplitude.
877: What is missing is a (vanishing for the classical group) factor of $\eta^2$ for each vertex of the triangulation.
878:
879: Note that, unlike the
880: case of 2D gravity where we have a clear interpretation of both the rank $N$ of the matrices
881: and the coupling constant $g$ (they are related to Newton and cosmological constant correspondingly),
882: in the case of Boulatov theory the interpretation of $g$ is obscure.
883: It is not anymore related to the volume of a tetrahedron, for the latter is now a function of the
884: spins. The appearance of Newton's constant is also not that direct. It serves to relate the
885: spin $l_e$ labeling edges to their physical (dimensionful) length. It has been argued recently \cite{Laurent-recent}
886: that the coupling constant $g$ should be thought of as the loop counting parameter that weights the
887: topology changing processes. However, this issue is not settled and we shall not comment on it any further.
888:
889: \subsection{Algebra structure}
890: \label{ss:algebra}
891:
892: Let us remind the reader that the algebra ${\mathcal A}^*$ of functions on
893: the group can be given a structure of the Hopf algebra. The reason why the algebra is referred to as
894: ${\mathcal A}^*$ and not $\mathcal A$ will become clear
895: below.\footnote{It is customary to denote the commutative point-wise
896: product of functions by $\bullet$ and the non-commutative
897: convolution product by $\star$, so we go against the convention
898: here. The reason for this unusual notation is that we decided to
899: embed $\mathcal A$ into the Drinfeld double $\D$ in a particular
900: way, see below, and this induces the $\bullet$ and $\star$ products
901: in the way chosen.} The Hopf algebra structure is as follows:
902: \be\nonumber
903: {\rm multiplication} &{}& (f_1\star f_2)(x)= f_1(x) f_2(x), \\ \nonumber
904: {\rm identity} &{}& i(x)= 1, \\ \nonumber
905: {\rm co-multiplication} &{}& (\Delta^* f)(x,y) = f(xy), \\
906: {\rm co-unit} &{}& \varepsilon(f) = f(e), \\ \nonumber
907: {\rm antipode} &{}& (\kappa f)(x)=f(x^{-1}), \\ \nonumber
908: {\rm involution} &{}& f^\circ(x) = \overline{f(x)}.
909: \ee
910: Of importance is also the Haar functional $h:{\mathcal A}\to \CC$:
911: \be
912: h^*(f) = \int dx\, f(x).
913: \ee
914: A non-degenerate pairing
915: \be
916: \langle f_1,f_2 \rangle = \int dx f_1(x) f_2(x)
917: \ee
918: identifies ${\mathcal A}^*$ with its dual $({\mathcal A}^*)^*={\mathcal A}$.
919:
920: The dual algebra $\mathcal A$ is also a Hopf algebra. A multiplication on $\mathcal A$ is denoted by $\bullet$ and is
921: introduced via:
922: \be
923: \langle f_1 \bullet f_2, g\rangle := \langle f_1 \otimes f_2, \Delta^* g\rangle.
924: \ee
925: One obtains:
926: \be\label{bullet}
927: (f_1\bullet f_2)(x) = \int dz\, f_1(z) f_2(z^{-1} x).
928: \ee
929: All other operations on $\mathcal A$ read:
930: \be
931: {\rm identity} &{}& i(x)= \delta_e(x), \\ \nonumber
932: {\rm co-multiplication} &{}& (\Delta f)(x,y) = f(x) \delta_x(y), \\
933: {\rm co-unit} &{}& \varepsilon(f) = \int dx f(x), \\ \nonumber
934: {\rm antipode} &{}& (\kappa f)(x)=f(x^{-1}), \\ \nonumber
935: {\rm involution} &{}& f^*(x) = \overline{f(x^{-1})}.
936: \ee
937: It is easy to check that $*$ is indeed an involution:
938: \be\label{inv}
939: (f_1\bullet f_2)^* = f_2^* \bullet f_1^*.
940: \ee
941: One also needs a Haar functional $h:{\mathcal A}\to \CC$:
942: \be
943: h(f)=f(e).
944: \ee
945:
946: Using this structure, a positive definite inner product can be
947: defined:
948: \be\label{ip}
949: \langle f_1, \overline{f_2} \rangle = \int dx f_1(x) \overline{f_2(x)}
950: = h(f_1 \bullet f_2^*)=h(f_2^* \bullet f_1).
951: \ee
952: The algebraic structure on $\mathcal A$ will play an important role when we
953: give an algebraic formulation of the Boulatov theory.
954:
955: \subsection{Projectors}
956:
957: Of special importance are functions on the group that satisfy the
958: projector property:
959: \be\label{proj-prop}
960: f\bullet f = f.
961: \ee
962: Examples of such projectors are given by:
963: (i) characters
964: \be\label{P-j}
965: P_j:={\rm dim}_j \chi_j(x),
966: \ee
967: and (ii) spherical functions ${\rm
968: dim}_j T^j_{00}(x)$. In both cases the projector property is readily
969: verified using \eqref{A-2-1}. We have:
970: \be
971: P_j \bullet P_k = \delta_{jk} P_j.
972: \ee
973: We will mostly be interested in these
974: character projectors. We note that the unit element, that is, the
975: $\delta$-function on the group
976: can be decomposed into the character projectors:
977: \be
978: i(x)=\delta_e(x)=\sum_j P_j.
979: \ee
980: Of importance for what follows is the following element of ${\mathcal A}^*\otimes{\mathcal A}^*$:
981: \be\label{G-prop}
982: G=(\kappa\otimes {\rm id})\Delta^* i, \\ \nonumber
983: G(x,y) = \delta_e(x^{-1} y).
984: \ee
985: In view of a property:
986: \be
987: \int dy\, G(x,y) G(y,z) = G(x,z)
988: \ee
989: this function on $G\times G$ can be referred to as a {\it propagator}. The propagator is just a kernel
990: of the operator $G$ acting on $\mathcal A$ via the $\bullet$-product.
991:
992: The unit element and the characters are the basic projectors on a
993: single copy of $\mathcal A$. More interesting projectors
994: can be constructed when working with several copies of the algebra.
995: Consider an element $\theta\in {\mathcal A}^{\otimes 3}$ obtained from
996: three propagators $G$:
997: \be
998: \theta:= (1^{\otimes 3} \otimes h_3) G\otimes G \otimes G.
999: \ee
1000: Here we have introduced a new operation $h_3:{\mathcal A}^* \otimes {\mathcal A}^* \otimes {\mathcal A}^* \to\CC$:
1001: \be
1002: h_3(f_1,f_2,f_3) = h^*(f_1\star f_2\star f_3) = \int dx f_1(x) f_2(x) f_3(x).
1003: \ee
1004: As a function on 3 copies of the group the element $\theta$ is given by:
1005: \be\label{theta}
1006: \theta(x_1,x_2,x_3) := \int dg\, \delta_e(x_1^{-1} g) \delta_e(x_2^{-1} g)
1007: \delta_e(x_3^{-1} g) =
1008: \sum_{j_i} \prod_i{\rm dim}_{j_i} \theta_{j_i}(x_i),
1009: \ee
1010: where
1011: \be\label{theta-j}
1012: \theta_{j_i}(x_i):= \theta_{j_1,j_2,j_3}(x_1,x_2,x_3):=\int dg\,\, \chi_{j_1}(x_1^{-1} g) \chi_{j_2}(x_2^{-1} g) \chi_{j_2}(x_2^{-1} g).
1013: \ee
1014: Let us note that $\theta$ is $*$-invariant: $\theta^*=\theta$.
1015: It is easy to verify that $\theta$ is a projector:
1016: \be
1017: \theta\bullet \theta = \theta.
1018: \ee
1019: To show this one uses \eqref{A-3} and \eqref{A-theta}. The function
1020: $\theta(x_1,x_2,x_3)$ projects onto functions invariant under the left
1021: diagonal action. Indeed, consider an arbitrary function
1022: $\phi(x_1,x_2,x_3)\in {\mathcal A}^{\otimes 3}$. Define
1023: the $\Phi$ to be $\theta$ with the projector $\theta$ applied on the
1024: left. The function $\Phi$ is invariant under the left diagonal shifts:
1025: \be
1026: \tilde{\phi}:=\theta\bullet \phi, \qquad \tilde{\phi}(g x_1,g x_2, g
1027: x_3)=\tilde{\phi}(x_1,x_2,x_3).
1028: \ee
1029: The field $\tilde{\phi}$ obtained this way is the basic field of Boulatov
1030: theory, see \eqref{sym}. The above more abstract formulation in
1031: algebra terms in necessary for a generalization to field theory
1032: on the Drinfeld double.
1033:
1034: \subsection{Algebraic formulation of Boulatov's theory}
1035:
1036: An equivalent formulation of Boulatov theory can be
1037: achieved using the Hopf algebra operations introduced
1038: earlier in this section. More specifically, we will need
1039: the second algebra structure in terms of the non-commutative
1040: $\bullet$-product and $*$-involution. Let us first write the
1041: kinetic term of the field theory action \eqref{GFT-1}. We have:
1042: \be\label{alg-kin}
1043: S_{kin}=\frac{1}{2}h((\theta\bullet\phi)^* \bullet (\theta\bullet\phi)) =
1044: \frac{1}{2}h(\phi^* \bullet\theta\bullet\phi).
1045: \ee
1046: Here we have used the fact \eqref{inv} that $*$ is an involution,
1047: and the fact that $\theta$ is a projector. All operations here, namely
1048: the $\bullet$-product, the $*$-involution, and the co-unit are
1049: understood in this formula as acting on ${\mathcal A}^{\otimes 3}$,
1050: separately on each of the copies of the algebra.
1051:
1052: There are two possible points of view on the kinetic term
1053: \eqref{alg-kin}. One is that the action is a functional of a field
1054: $\tilde{\phi}$ that satisfies the invariance property
1055: $\theta\bullet\tilde{\phi}=\tilde{\phi}$. This is the point of view
1056: taken in the original formulation of Boulatov theory that has been
1057: described above. The other interpretation is suggested by
1058: \eqref{alg-kin}, and is to view $\theta$ in this expression as the
1059: kinetic term ``differential operator'', or the inverse of the
1060: propagator of the theory. The action is then invariant under the
1061: following symmetry: $\delta\phi=\psi$, where $\psi$ is a field
1062: satisfying $\theta\bullet\psi=0$. This symmetry should be viewed as
1063: gauge symmetry of the theory; the ``physical'' degrees of freedom are
1064: not those of $\phi$ but those of the $\theta$-cohomology classes.
1065: Boulatov theory in its original formulation \eqref{GFT-1}
1066: can be viewed as the gauge-fixed version of this theory. The gauge symmetry
1067: described is of great importance. For example, one can consider a theory
1068: with no $\theta$-projector inserted in the
1069: action. This theory does not have any gravitational interpretation, as it is
1070: easy to check. Thus, it is the presence of $\theta$ in the action, and thus
1071: the extra gauge symmetry that ensures a relation to gravity. It can be
1072: argued that this symmetry is the usual diffeomorphism-invariance
1073: of a gravitational theory in disguise.
1074:
1075: Once the kinetic term is understood, it is easy to write down the
1076: interaction term as well. One has to take 4 fields $\theta\bullet\phi$
1077: and $\bullet$-multiply them all as is suggested by the structure of the
1078: interaction term in \eqref{GFT-1} to obtain an element of
1079: ${\mathcal A}^{\otimes 6}$. Each copy of the algebra is a product of
1080: two fields; the $*$-involution has to taken on one of them. After
1081: all fields are multiplied, the Haar functional $h$ has to be applied to get a
1082: number. We will not write the corresponding expression as it is rather
1083: cumbersome. The structure arising is best understood using the language of operator kernels
1084: to which we now turn.
1085:
1086: \subsection{Formulation in terms of kernels}
1087:
1088: We have given an algebraic formulation of Boulatov theory, in which fields are viewed
1089: as operators in ${\mathcal A}^{\otimes 3}$, and one uses the
1090: $\bullet$-product on $\mathcal A$ to write the action. An equivalent formulation can be given by introducing
1091: kernels of all the operators. Such a formulation is more familiar in
1092: a field theory context, and
1093: will be quite instrumental in dealing with the theory.
1094:
1095: Let us interpret the formula \eqref{bullet} as follows. The quantity:
1096: \be
1097: f_2(x,y):=f_2(x^{-1} y) = (\kappa\otimes {\rm id})\Delta^* f_2,
1098: \ee
1099: which is an element of ${\mathcal A}^*\otimes {\mathcal A}^*$,
1100: is interpreted as the kernel of an operator ${\mathcal O}_{f_2}$ corresponding to $f_2$ that acts on $f_1$ from
1101: the right:
1102: \be
1103: f_1 \circ {\mathcal O}_{f_2} = \int dz f_1(z) f_2(z,x).
1104: \ee
1105: Thus, instead of dealing with operators from $\mathcal A$ one can work with kernels from
1106: ${\mathcal A}^*\otimes {\mathcal A}^*$.
1107:
1108: The inner product \eqref{ip} can also be expressed in terms of the kernels. Since
1109: $f(y,z):=f(y^{-1} z)$, the kernel for $f^*$ is given by:
1110: \be
1111: f^*(y,z)=\overline{f(y z^{-1})}.
1112: \ee
1113: Therefore:
1114: \be
1115: \langle f_1, \overline{f_2}\rangle = \int dx dy \,\, f_1(x,y) f_2^*(y,x).
1116: \ee
1117: Here we have taken a convolution of the kernels of two operators, and then took the
1118: Haar functional given by the trace of the corresponding kernel:
1119: \be
1120: h(f) \to \int dx f(x,x).
1121: \ee
1122:
1123: Let us now introduce kernels for all the objects that are necessary to define Boulatov theory. The
1124: kernel for the field $\phi\in {\mathcal A}^*\otimes {\mathcal A}^*\otimes {\mathcal A}^*$ is:
1125: \be\label{ker-phi}
1126: \phi := ((\kappa\otimes {\rm id})\Delta^*)^{\otimes 3} \phi, \\ \nonumber
1127: \phi(x_1,y_1;x_2,y_2;x_3,y_3):= \phi(x_1^{-1} y_1,x_2^{-1} y_2, x_3^{-1} y_3).
1128: \ee
1129: The kernel for the projector $\theta$ is similarly given by:
1130: \be\label{ker-theta}
1131: \theta(x_1,y_1;x_2,y_2;x_3,y_3):= \theta(x_1^{-1} y_1,x_2^{-1} y_2, x_3^{-1} y_3).
1132: \ee
1133: The action of $\theta$ on the field now reads:
1134: \be
1135: &{}&\Phi(x_1,y_1;x_2,y_2;x_3,y_3):= \\ \nonumber
1136: &{}& \int dz_1 dz_2 dz_3 \, \phi(x_1,z_1;x_2,z_2;x_3,z_3)
1137: \theta(z_1,y_1;z_2,y_2;z_3,y_3).
1138: \ee
1139:
1140: It is now easy to write the action for the theory. We shall assume that the
1141: field is real. The action reads:
1142: \be\label{boul-kern}
1143: S= \frac{1}{2} \int \prod_{i=1}^3 dx_i dy_i dz_i\,\, \phi(x_i,y_i) \theta(y_i,z_i) \phi(z_i,x_i) +
1144: \\ \nonumber
1145: \frac{g}{4!} \int \prod_{i\not=j=1}^{4} dx_{ij} dy_{ij} dz_{ij} \,\,
1146: \phi(x_{ij},y_{ij}) \theta(y_{ij},z_{ij}) \theta(z_{ji},y_{ji}) \phi(y_{ji},x_{ji}).
1147: \ee
1148: Here $x_{ij}=x_{ji}, z_{ij}=z_{ji}$, but $y_{ij}\not=y_{ji}$ and are independent
1149: integration variables. We have also introduced a notation: $\phi(x_i,y_i):=\phi(x_1,y_1;x_2,y_2;x_3,y_3)$
1150: and $\phi(x_{ij},y_{ij}):=\phi(x_{ij},y_{ij};x_{ik},y_{ik};x_{il},y_{il})$ with
1151: $i\not=j\not=k\not=l$.
1152:
1153: \subsection{Remarks on Boulatov theory}
1154:
1155: Let us remark on possible choices of propagators in
1156: \eqref{boul-kern}. One can try to use a different operator in place of $\theta$ in both
1157: the kinetic and the potential terms. Or, more generally, one can have an
1158: interaction term that is built of quantities different than the one used
1159: in the kinetic term, see \cite{Rovelli} where such more general theories
1160: are considered. However, this more general class of theories fails to
1161: lead to 3-manifold invariants. Related to this is the fact that these
1162: more general theories do not have a dual Feynman diagram
1163: interpretation. As we have discussed in the introduction, we would
1164: like to restrict our attention to a special class of theories, namely
1165: those in which Feynman amplitudes that follow from the group field
1166: theory expansion also have an interpretation in terms of Feynman
1167: diagrams for the dual complex. In other words, as we have discussed,
1168: group field theory Feynman diagrams have the interpretation of
1169: amplitudes for a complex dual to some triangulated 3-manifold. The
1170: triangulation itself can be considered as a Feynman diagram. When can
1171: the group field theory amplitude, which is the one for the dual triangulation,
1172: be interpreted as a Feynman amplitude for the original triangulation
1173: as well? Thus, the question we are posing is which group field theories
1174: admit a dual formulation. As is clear from our discussion of Boulatov and
1175: Ponzano-Regge models in the introduction, this particular theory does admit
1176: both interpretations. Are there any other theories with a similar property?
1177:
1178: We shall not attempt to answer this question in its full generality. We shall only
1179: make a remark concerning the choice of the propagator of the model. As we have seen,
1180: the propagator of any 3d group field theory model should consist of 3 strands. Thus,
1181: it can be described as made out of 3 propagators, one for each strand. This is
1182: what happens in the case of Boulatov model, where the $\theta$-projector is made
1183: out of 3 $G$-propagators \eqref{G-prop}. Now each strand of the group field
1184: theory diagram that forms a closed loop is interpreted as dual to an
1185: edge of a triangulated 3-manifold.
1186: When this triangulation is itself interpreted as a Feynman diagram, there
1187: must be a propagator for every edge. It is clear that this propagator will be
1188: built from the group field theory strand propagator $G$, and will just be
1189: a certain power of it, with the power given by a number of dual edges
1190: forming the boundary of the dual face. Because different triangulations
1191: can have this number different, to have the interpretation we are after
1192: the propagator $G$ must be a projector $G\bullet G=G$. Only such
1193: propagators admit the dual Feynman diagram interpretation. This requirement
1194: is clearly very restrictive, and limits the choice of possible propagators
1195: dramatically. In this sense the models of the type we are considering in this paper
1196: are very scarce.
1197:
1198: Now that we have understood how to formulate Boulatov's theory in abstract terms, let us
1199: construct an analogous theory with the group manifold replaced by the Drinfeld double of
1200: $\SU(2)$.
1201:
1202: \section{Drinfeld double of $\SU(2)$}
1203: \label{sec:DD}
1204:
1205: The Drinfeld double of a classical group was first studied in \cite{Philippe}.
1206: Our description of the Drinfeld double $D(G)$ of $G=\SU(2)$ closely
1207: follows that in \cite{Koorn}. Following this reference, we describe
1208: $D(G)$ as the space of functions on it. As a linear space $D(G)$ is identified with the space
1209: $C(G\times G)$ of functions on two copies of the group. On $D(G)$ we have a non-degenerate pairing:
1210: \be
1211: \langle f_1,f_2 \rangle := \int dx dy\,\, f_1(x,y) f_2(x,y).
1212: \ee
1213: This pairing identifies the dual $D(G)^\star$ of the Drinfeld double with $C(G\times G)$.
1214: The following operations are defined on $D(G)$:
1215: \be
1216: \nonumber
1217: {\rm multiplication} &{}& (f_1\bullet f_2)(x,y)= \int dz\, f_1(x,z) f_2(z^{-1}xz,z^{-1}y), \\ \nonumber
1218: {\rm identity} &{}& 1(x,y)= \delta_e(y), \\ \nonumber
1219: {\rm co-multiplication} &{}& (\Delta f)(x_1,y_1;x_2,y_2) =
1220: f(x_1x_2,y_1)\delta_{y_1}(y_2), \\ \label{dd}
1221: {\rm co-unit} &{}& \varepsilon(f) = \int dy\, f(e,y), \\ \nonumber
1222: {\rm antipode} &{}& (\kappa f)(x,y)=f(y^{-1} x^{-1} y,y^{-1}), \\ \nonumber
1223: {\rm involution} &{}& f^\ast(x,y) = \overline{f(y^{-1} xy,y^{-1})}.
1224: \ee
1225:
1226: By duality we have the following operations on the dual $D(G)^*$:
1227: \be
1228: \nonumber
1229: {\rm multiplication} &{}& (f_1\star f_2)(x,y)= \int dz\, f_1(z,y) f_2(z^{-1}x,y), \\ \nonumber
1230: {\rm identity} &{}& i(x,y)= \delta_e(x), \\ \nonumber
1231: {\rm co-multiplication} &{}& (\Delta^\star f)(x_1,y_1;x_2,y_2) =
1232: f(x_1,y_1y_2)\delta_{x_2}(y_1^{-1}x_1y_1), \\ \label{dd-dual}
1233: {\rm co-unit} &{}& \varepsilon^\star(f) = \int dx\, f(x,e), \\ \nonumber
1234: {\rm antipode} &{}& (\kappa^\star f)(x,y)=f(y^{-1} x^{-1} y,y^{-1}), \\ \nonumber
1235: {\rm involution} &{}& f^\circ(x,y) = \overline{f(x^{-1},y)}.
1236: \ee
1237:
1238: The universal R-matrix $R\in D(G)\otimes D(G)$ is given by:
1239: \be\label{R-matr}
1240: R(x_1,y_1;x_2,y_2)=\delta_e(y_1) \delta_e(x_1 y_2^{-1}).
1241: \ee
1242: We will also need the central ribbon element:
1243: \be
1244: c(x,y)=\delta_e(xy)
1245: \ee
1246: and the monodromy element:
1247: \be
1248: Q(x_1,y_1;x_2,y_2):=(R\bullet \sigma R)(x_1,y_1;x_2,y_2)=\delta_{y_1}(x_2) \delta_{y_2}(x_2^{-1}x_1 x_2).
1249: \ee
1250: Here $\sigma(x_1,y_1;x_2,y_2)=(x_2,y_2;x_1,y_1)$.
1251:
1252: The Haar functionals $h:D(G)\to \CC$ and $h^\star:D(G)^*\to\CC$ are given by:
1253: \be
1254: h(f)=\int dx\, f(x,e), \qquad h^\star(f)=\int dy\, f(e,y).
1255: \ee
1256: Note that the Haar functional $h^*$ on $D(G)^*$ coincides with the
1257: co-unit $\epsilon$ on $D(G)$, and similarly $h$ coincides with $\epsilon^*$.
1258: Using these functionals, a positive-definite inner product on $D(G)$ can be
1259: defined as:
1260: \be
1261: \langle f_1, \overline{f_2} \rangle = \int dx dy\, f_1(x,y)
1262: \overline{f_2(x,y)} = h(f_1\bullet f_2^*) = h^*(f_1 \star f_2^\circ).
1263: \ee
1264:
1265: To acquire a better understanding of the Drinfeld double, let us
1266: consider some of its sub-algebras.
1267:
1268: \subsection{Sub-algebras of the Drinfeld double}
1269:
1270: The Drinfeld double is a twisted product of the algebra of functions on the
1271: group and the group itself.
1272: The subalgebra ${\mathcal A}^*$ of functions on the group described in the previous
1273: section is represented by the elements:
1274: \be
1275: f(g) \to \delta_e(x) f(y) \in D(G).
1276: \ee
1277: Then it is easy to check that the $\star$-product coincides with the usual point-wise
1278: multiplication of functions:
1279: \be
1280: (\delta_e f_1 \star \delta_e f_2)(x,y)=\delta_e(x) f_1(y) f_2(y).
1281: \ee
1282: It is therefore clear that the $\star$-algebraic structure on the algebra ${\mathcal A}^*$ of functions
1283: on the group that we have described in section \ref{sec:Boul} is
1284: exactly the one obtained from the Drinfeld double dual algebra structure
1285: \eqref{dd-dual} when specialized to functions of the form $\delta_e(x) f(y)$.
1286: Note that the algebra ${\mathcal A}^*$
1287: with its pointwise multiplication can also be embedded into the Drinfeld algebra itself
1288: with its $\bullet$-product. Indeed, as is easy to check, the $\bullet$-product on
1289: functions of the form $f(x) \delta_e(y)$ is just the usual pointwise multiplication.
1290:
1291: The group is represented by elements of the form:
1292: \be
1293: g\to \delta_g(y) \in D(G).
1294: \ee
1295: The $\bullet$-multiplication reduces to the usual group multiplication law:
1296: \be
1297: (\delta_{g_1} \bullet \delta_{g_2})(x,y) = \delta_{g_1g_2}(y).
1298: \ee
1299:
1300: \subsection{Irreducible representations}
1301:
1302: Let us denote by $C_\theta, \theta\in[0,\pi]$ the conjugacy classes in
1303: $\SU(2)$, and by $g_\theta$ a representative of $C_\theta$ that is
1304: in the Cartan subgroup $U(1)$. The irreducible representations are
1305: labeled by pairs $(\theta,k)$ of conjugacy classes and
1306: representations of the centralizer $U(1)$. The carrier space $V^{(\theta,k)}$ is:
1307: \be
1308: V^{(\theta,s)} := \{ \phi:G\to \CC | \phi(x h) = e^{-i s\xi} \phi(x),
1309: \forall x\in G, h={\rm diag}(e^{i\xi},e^{-i\xi}) \}
1310: \ee
1311: and the action of an element $f\in D(G)$ is:
1312: \be
1313: (\pi^{(\theta,s)}(f) \phi)(x) = \int dz\, f(x g_\theta
1314: x^{-1},z)\phi(z^{-1} x).
1315: \ee
1316: An orthonormal basis in $V^{(\theta,s)}$ is given by the matrix elements:
1317: $\sqrt{{\rm dim}_j} T^j_{ms}(x), j\geq s, -j\leq m\leq j$. We shall
1318: also use the bra-ket notation and denote the basis vectors by
1319: $|jm\rangle$. The carrier space $V^{(\theta,s)}$ can be decomposed
1320: into finite dimensional subspaces:
1321: \be
1322: V^{(\theta,s)} = \oplus_{j\geq s} V^{(\theta,s)}_j, \qquad
1323: V^{(\theta,s)}_j = {\rm Span}\{ |jm\rangle \}.
1324: \ee
1325:
1326: \subsection{Matrix elements and characters}
1327:
1328: Let us consider the matrix elements:
1329: \be\label{mm-1}
1330: (\pi^{(\theta,s)}(f))^{jj'}_{mm'} = \langle jm| \pi^{(\theta,s)}(f)
1331: |j'm'\rangle = \\ \nonumber \sqrt{{\rm dim}_j {\rm dim}_{j'}} \int dx dz\,
1332: \overline{T^j_{ms}(y)} f(y g_\theta y^{-1}, z) T^{j'}_{m's}(z^{-1}y).
1333: \ee
1334: By definition, the matrix elements are in $D(G)^*$. Using the pairing
1335: $\langle\cdot,\cdot\rangle$ we can identify them with functions on
1336: $G\times G$. To this end, let us transform the above formula. Let us
1337: introduce, for each element $x\in C_\theta$ of the conjugacy class $C_\theta$, an
1338: element $B_x\in G$ such that:
1339: \be
1340: B_x g_\theta B_x^{-1}=x.
1341: \ee
1342: We will also need a $\delta$-function $\delta_\theta(x)$ picked on a conjugacy class $\theta$, which is
1343: defined as:
1344: \be\label{delta-t}
1345: \int dg\, \delta_\theta(g) f(g) = \int_{G/H} f(x g_\theta x^{-1}) dx.
1346: \ee
1347: Using these objects the matrix element can be written as:
1348: \be\label{mm-2}
1349: (\pi^{(\theta,s)}(f))^{jj'}_{mm'} = \sqrt{{\rm dim}_j {\rm dim}_{j'}} \int dx dy\,
1350: f(x, y) \delta_\theta(x) \overline{T^j_{ms}(B_x)} T^{j'}_{m's}(y^{-1} B_x).
1351: \ee
1352: Therefore we define:
1353: \be\label{me}
1354: (\pi^{(\theta,s)})^{jj'}_{mm'}(x,y) := \sqrt{{\rm dim}_j {\rm dim}_{j'}}
1355: \delta_\theta(x) \overline{T^j_{ms}(B_x)} T^{j'}_{m's}(y^{-1} B_x).
1356: \ee
1357: This is the main formula for matrix elements as functions on $G\times G$.
1358:
1359: The character is obtained in the usual way as the trace:
1360: \be\label{char}
1361: \chi^{(\theta,s)}(x,y) := \sum_{jm} (\pi^{(\theta,s)})^{jj}_{mm}(x,y) =
1362: \delta_\theta(x) \sum_j {\rm dim}_j T^{j}_{ss}(B_x^{-1} y^{-1} B_x).
1363: \ee
1364: To simplify this further let us use the fact that:
1365: \be
1366: \sum_j {\rm dim}_j T^j_{ss}(g) = e^{-is(\phi+\psi)}
1367: \delta(\cos{\theta}).
1368: \ee
1369: We have used the Euler parameterization here:
1370: $g=g(\phi,\theta,\psi)$. Applied to \eqref{char} this formula implies that
1371: $y=B_x g_\xi B_x^{-1}$, or, in other words, that \eqref{char} contains
1372: the $\delta$-function $\delta_{xy}(yx)$. Thus, we have:
1373: \be
1374: \chi^{(\theta,s)}(x,y) = \delta_\theta(x) \delta_{xy}(yx)
1375: \chi_s(B_x^{-1} y^{-1} B_x),
1376: \ee
1377: where $\chi_s(n)=e^{is\xi}, n={\rm diag}(e^{i\xi},e^{-i\xi})$. This is
1378: the character formula given in \cite{Koorn}.
1379:
1380: \subsection{Decomposition of the identity}
1381: Now that we have obtained the characters, we can use them to decompose
1382: the identity element in the Drinfeld double. To this end, we will need
1383: the Weyl integration formula:
1384: \be
1385: \int_G dg\, f(g) = \int_{H/W} \frac{d\theta}{\pi} \, \Delta(\theta)^2 \int_{G/H} dx\, f(x
1386: g_\theta x^{-1}).
1387: \ee
1388: Together with \eqref{delta-t} this implies:
1389: \be\label{ident-1}
1390: \int_{H/W} \frac{d\theta}{\pi} \, \Delta(\theta)^2 \delta_\theta(g) =
1391: 1.
1392: \ee
1393: Let us now apply the summation over $s$ to \eqref{char}. We get:
1394: \be\label{1-decomp}
1395: \int_{H/W} \frac{d\theta}{\pi} \, \Delta(\theta)^2 \sum_s \chi^{(\theta,s)}(x,y)=\delta_e(y).
1396: \ee
1397:
1398: \section{Group field theory for the Drinfeld double}
1399: \label{sec:gft}
1400:
1401: From the algebraic formulation of Boulatov's theory in section \ref{sec:Boul} it was clear that
1402: the main object that is used in the construction of the model is the projector
1403: $\theta\in{\mathcal A}^{\otimes 3}$. Unlike the case of Boulatov theory,
1404: where there is essentially a unique such projector constructed from
1405: the identity element of $\mathcal A$, in the Drinfeld double case
1406: there are several interesting ``identity elements'' that can be used
1407: in the construction of $\theta$. We shall analyze several
1408: different possibilities.
1409:
1410: An explanation of why different possibilities can arise is as
1411: follows. The $\theta$-projector that defines GFT should be constructed from a
1412: projector on $\D$. Characters of irreducible representations are
1413: projectors. To construct a more general projector one can take various linear combinations of
1414: characters. Natural examples are given by: (i) the sum of
1415: characters of all the representations of $\D$, which gives the
1416: identity operator; (ii) the sum of characters of all simple
1417: representations; (iii) the character of the trivial
1418: representation. One can more generally consider not characters but
1419: individual matrix elements, which are also projectors. In this paper we will
1420: consider the simplest case of the projector constructed from the
1421: identity operator, as well as another in a certain sense dual projector constructed from the
1422: element $\delta_e(x)$.
1423:
1424: \subsection{Projector $\theta$ constructed from the identity $\delta_e(y)$}
1425:
1426: Let us first consider a direct analog of Boulatov model. Thus, we
1427: construct a projector $\theta$ from the identity element
1428: $\delta_e(y)$ on $\mathcal D$.
1429: To construct the projector we follow the same procedure that was used in section \ref{sec:Boul}. Namely,
1430: let us first construct the propagator (or the kernel of the identity
1431: operator $\delta_e(y)$):
1432: \be
1433: {\mathcal D}^* \otimes {\mathcal D}^* \ni G = (\kappa\otimes{\rm id})\Delta^* 1, \\ \nonumber
1434: G(x_1,y_1;x_2,y_2) = \delta_e(x_1 x_2) \delta_e(y_1 y_2^{-1}).
1435: \ee
1436: Let us define an operator
1437: $h_3:{\mathcal D}^* \otimes {\mathcal D}^* \otimes {\mathcal D}^* \to \CC$ via:
1438: \be\label{h3}
1439: &{}& h_3(f_1,f_2,f_3):=h^*(f_1\star f_2\star f_3) = \\ \nonumber
1440: &{}&\int dz_1 dz_2 dz_3 dy \,\, f_1(z_1,y)
1441: f_2(z_2,y) f_3(z_3,y) \delta_e(z_1 z_2 z_3).
1442: \ee
1443: The $\theta$-projector is obtained by applying $h_3$ to $G^{\otimes 3}$:
1444: \be
1445: \theta:=({\rm id}^{\otimes 3}\otimes h_3)\, G\otimes G\otimes G.
1446: \ee
1447: Let us find an explicit expression for this projector. We have:
1448: \be\label{theta*}
1449: \theta(x_1,y_1;x_2,y_2;x_3,y_3) =
1450: \int dz_1 dz_2 dz_3 dy \,\, \delta_e(x_1 z_1) \delta_e(x_2 z_2) \delta_e(x_3 z_3) \times \\ \nonumber
1451: \delta_e(z_1 z_2 z_3) \delta_e(y_1^{-1} y)\delta_e(y_2^{-1} y)\delta_e(y_3^{-1} y)=
1452: \delta_e(x_1^{-1} x_2^{-1} x_3^{-1}) \theta(y_1,y_2,y_3),
1453: \ee
1454: where $\theta(y_1,y_2,y_3)$ is given by \eqref{theta}. An explicit verification shows that
1455: $\theta$ is real $\theta^*=\theta$. We also have to check the
1456: projector property of $\theta$. However, unlike the case of Boulatov
1457: model, the quantity \eqref{theta*} is not a projector. The
1458: $\bullet$-product of two $\theta$ gives a divergent factor of
1459: $\eta^{-2}$ \eqref{eta} that we have already encountered before in the discussion
1460: of the Ponzano-Regge model. We nevertheless proceed formally, and
1461: introduce a multiple of $\theta$ as a projector. The
1462: following lemma is a statement to this effect.
1463: \begin{lemma}The quantity $\eta^2\theta$, where
1464: $\eta^{-2}=\delta_e(e)$ is a projector:
1465: \be
1466: (\eta^2 \theta)\bullet (\eta^2\theta) = \eta^2 \theta.
1467: \ee
1468: \begin{proof}
1469: A proof is by verification. We have:
1470: \be
1471: (\theta\bullet\theta)(x_i,y_i)=\delta_e(x_1^{-1}x_2^{-1}x_3^{-1}) \\
1472: \nonumber \int
1473: dg\, \prod_i \int dz_i\,\delta_g(z_i) \int d\tilde{g}\, \delta_e(z_1^{-1}
1474: x_1^{-1} z_1 z_2^{-1} x_2^{-1} z_2 z_1^{-1} x_1^{-1} z_1) \prod_i
1475: \delta_{\tilde{g}}(z_i^{-1} y_i) = \\ \nonumber
1476: (\delta_e(x_1^{-1} x_2^{-1} x_3^{-1}))^2 \int dg d\tilde{g}\, \prod_i
1477: \delta_{\tilde{g}}(g^{-1} y_i) = \eta^{-2} \theta(x_i,y_i).
1478: \ee
1479: \end{proof}
1480: \end{lemma}
1481:
1482: The ``projector'' so defined is the weakest point of our
1483: construction. A sceptic may argue that this quantity does not exist or
1484: is zero. We note, however, that both such a projector and the $\eta$
1485: factor would be perfectly
1486: well-defined had we been working with the Drinfeld double of the
1487: quantum $\SU_q(2)$ at root of unity instead. Passing to the quantum
1488: $\SU_q(2)$ is known to have the interpretation of adding a
1489: (positive) cosmological constant. Thus, as we see, unlike the Boulatov
1490: model itself, the model for the Drinfeld double of $\SU(2)$ is defined
1491: only formally. Later we shall consider another model which is
1492: well-defined.
1493:
1494:
1495: The quantity $\theta$ can be expressed in terms of the $R$-matrix. To state a result to this effect, we
1496: define the following graphical notation:
1497: \be
1498: R(\phi\otimes\psi)(x_1,y_1;x_2,y_2)=\phi(x_1,y_1)\psi(x_1^{-1}x_2 x_1,x_1^{-1} y_2) =\,\,
1499: \lower0.4in\hbox{\epsfig{figure=r.eps, height=0.8in}}
1500: \ee
1501: The opposite braiding is given by:
1502: \be
1503: \sigma R(\phi\otimes\psi)(x_1,y_1;x_2,y_2)=\phi(x_2^{-1} x_1 x_2, x_2^{-1} y_1)\psi(x_2,y_2) =\,\,
1504: \lower0.4in\hbox{\epsfig{figure=sr.eps, height=0.8in}}
1505: \ee
1506: The following result is confirmed by an explicit computation:
1507: \begin{lemma}
1508: \be\label{r-lemma}
1509: \lower0.3in\hbox{\epsfig{figure=braid.eps, height=0.8in}} \,\,=
1510: \phi(x^{-1} x_1^{-1}\ldots x_n^{-1}x x_n \ldots x_1 x,
1511: x^{-1} x_1^{-1} \ldots x_n^{-1}xy) \times \\ \nonumber
1512: \prod_{i=1}^n \psi_i(x^{-1} x_i x, x^{-1} y_i).
1513: \ee
1514: \end{lemma}
1515: Let us now take $\phi(x,y)=\delta_e(y)$, and take the Haar functional
1516: $h$ in the first channel. We get the following result:
1517: \begin{lemma}
1518: \be\label{general-braid}
1519: \lower0.3in\hbox{\epsfig{figure=gen-braid.eps, height=0.8in}} \,\,=
1520: \delta_e(x_1^{-1}\ldots x_n^{-1}) \int dg \prod_i \psi_i(g^{-1} x_i
1521: g,g^{-1} y_i).
1522: \ee
1523: \end{lemma}
1524: We can use this result to give another description of the quantity $\theta$:
1525: \begin{lemma} \label{lemma-1}
1526: \be
1527: \lower0.25in\hbox{\epsfig{figure=c-braid.eps, height=0.8in}} \,\,= \theta
1528: \ee
1529: \end{lemma}
1530:
1531: Let us consider the object \eqref{general-braid}.
1532: We will need two properties of such objects, which are
1533: referred to as the {\it handleslide} and {\it killing}:
1534: \begin{lemma} A composition of two quantities \eqref{general-braid}
1535: satisfies the following ``handleslide'' property:
1536: \be\label{handleslide}\label{lemma-2}
1537: \lower0.4in\hbox{\epsfig{figure=handleslide-1.eps, height=0.8in}}=
1538: \lower0.55in\hbox{\epsfig{figure=handleslide-2.eps, height=1.1in}}
1539: \ee
1540: \begin{proof}A direct computation of the left hand side gives:
1541: \be\label{handle-1}
1542: \delta_e(x_1^{-1} \ldots x_k^{-1}) \int dg \, \prod_{i=1}^k
1543: \psi_i(g^{-1} x_i g,g^{-1} y_i) \times \\ \nonumber
1544: \delta_e(x_{k+1}^{-1} \ldots x_n^{-1}) \int d\tilde{g} \, \prod_{i=k+1}^n
1545: \psi_i(\tilde{g}^{-1} x_i \tilde{g},\tilde{g}^{-1} y_i).
1546: \ee
1547: The right-hand side is given by:
1548: \be\label{handle-2}
1549: \delta_e(x_1^{-1}\ldots x_n^{-1}) \delta_e(x_1^{-1}\ldots x_k^{-1}) \\
1550: \nonumber
1551: \int dgd\tilde{g} \, \prod_{i=1}^k \psi_i(g^{-1} \tilde{g}^{-1} x_i
1552: \tilde{g} g, g^{-1} \tilde{g}^{-1} y_i) \prod_{i=k+1}^n \psi_i(g^{-1}
1553: x_i g, g^{-1} y_i).
1554: \ee
1555: A straightforward change of integration variable in the first part of
1556: the product in \eqref{handle-2} makes it identical to \eqref{handle-1}
1557: and proves the lemma.
1558: \end{proof}
1559: \end{lemma}
1560:
1561: \begin{lemma} \label{lemma-3} The following killing property holds:
1562: \be\label{killing}
1563: \lower0.3in\hbox{\epsfig{figure=killing-1.eps, height=0.8in}}=
1564: \lower0.3in\hbox{\epsfig{figure=killing-2.eps, height=0.8in}}
1565: \ee
1566: \begin{proof} We take $\psi_n(x_n,y_n)=\delta_e(y_n)$,
1567: compute \eqref{general-braid}, and apply the Haar functional $h$ in
1568: the last channel. We get:
1569: \be
1570: \int dx_n \delta_e(x_1^{-1} \ldots x_n^{-1}) \int dg \,
1571: \prod_{i=1}^{n-1} \psi_i(g^{-1} x_i g,g^{-1}y_i) \delta_e(g^{-1})=
1572: \prod_{i=1}^{n-1} \psi_i(x_i,y_i).
1573: \ee
1574: This proves the lemma.
1575: \end{proof}
1576: \end{lemma}
1577:
1578: For completeness, let us also give a decomposition of the quantity $\theta$ into
1579: characters:
1580: \begin{lemma} The following decomposition of the quantity $\theta$
1581: holds:
1582: \be
1583: &{}&\theta(x_1,y_1;x_2,y_2;x_3,y_3)= \int_{H/W} \prod_{i=1}^3 \frac{d\theta_i}{\pi} \, \Delta(\theta_i)^2 \times
1584: \\ \nonumber
1585: &{}& \sum_{j_1,j_2,j_3} {\rm dim}_{j_1} {\rm dim}_{j_2} {\rm dim}_{j_3}
1586: \theta^{\theta_1,\theta_2,\theta_3}_{j_1,j_2,j_3}(x_1,y_1;x_2,y_2;x_3,y_3),
1587: \ee
1588: where
1589: \be\label{theta-char}
1590: &{}&\theta^{\theta_1,\theta_2,\theta_3}_{j_1,j_2,j_3}(x_1,y_1;x_2,y_2;x_3,y_3)= \\ \nonumber
1591: &{}&\delta_{\theta_1}(x_1) \delta_{\theta_2}(x_2) \delta_{\theta_3}(x_3)
1592: \delta_e(x_1^{-1} x_2^{-1} x_3^{-1}) \theta_{j_1,j_2,j_3}(y_1,y_2,y_3),
1593: \ee
1594: where $\theta_{j_i}(y_i)$ is given by \eqref{theta-j}.
1595: \begin{proof}
1596: Let us use the character decomposition of the identity operator $\delta_e(y)$:
1597: \be
1598: \delta_e(y) = \int_{H/W} \frac{d\theta}{\pi} \Delta(\theta)^2 \sum_j
1599: {\rm dim}_j \,\chi^\theta_j(x,y),
1600: \ee
1601: where
1602: \be\label{char*}
1603: \chi^{\theta}_j(x,y) := \sum_{s,m} (\pi^{(\theta,s)})^{jj}_{mm}(x,y) =
1604: \delta_\theta(x) \chi_j(y).
1605: \ee
1606: We can now apply the formula \eqref{general-braid} to a product of
1607: 3 characters, which results in \eqref{theta-char}.
1608: \end{proof}
1609: \end{lemma}
1610:
1611: \subsection{Projector constructed from $\delta_e(x)$}
1612:
1613: Another projector that we consider in this paper will be built from
1614: the operator $\hat{1}(x,y)=\delta_e(x) \in\D$. This element is
1615: {\it not} an identity: $(\hat{1} \bullet f)(x,y)= \delta_e(x) h^*(f)$.
1616: As we see, it plays the role of the element dual to the Haar measure
1617: on $\D^*$. We have put a hat above the symbol
1618: denoting this operator in order not to confuse it with the identity
1619: $\delta_e(y)$ in the Drinfeld double. It is clear that the
1620: $\theta$-projector constructed from $\hat{1}$ according to the rules
1621: as described above is equal to $\eta^{-2} \hat{1}^{\otimes 3}$. Indeed, we take
1622: 3 operators $\hat{1}$ and enclose them with a strand with $\hat{1}$
1623: inserted. The result \eqref{r-lemma} proves the assertion. It is
1624: obvious that a multiple of $\theta$ constructed this way is (formally) a
1625: projector. Another important property of the operator $\hat{1}$ is that
1626: it satisfies the handleslide and killing properties. Let us state two lemmas to this effect.
1627: \begin{lemma} \label{lemma-h} When the projector $P=\hat{1}$ is inserted
1628: into the meridian link, the handleslide property holds.
1629: \begin{proof}
1630: The left hand side is given by:
1631: \be
1632: \delta_e( x_1^{-1} \ldots x_k^{-1}) \prod_{i=1}^k \psi_i(x_i, y_i)
1633: \delta_e( x_{k+1}^{-1} \ldots x_n^{-1}) \prod_{i=k+1}^{n} \psi_i(x_i,y_i).
1634: \ee
1635: The right-hand side is given by:
1636: \be
1637: \delta_e( x_1^{-1} \ldots x_k^{-1} x_{k+1}^{-1} \ldots x_n^{-1}) \prod_{i=1}^k \psi_i(x_i, y_i)
1638: \delta_e( x_{k+1}^{-1} \ldots x_n^{-1}) \prod_{i=k+1}^{n} \psi_i(x_i,y_i).
1639: \ee
1640: The handleslide property is obvious.
1641: \end{proof}
1642: \end{lemma}
1643:
1644: \begin{lemma} \label{killing-h} The quantity $\hat{1}$ satisfies the
1645: killing property provided the operator inserted into the longitude
1646: is a function of $y$ only.
1647: \begin{proof} A proof is by direct verification.
1648: \end{proof}
1649: \end{lemma}
1650:
1651: Having discussed several different projectors that can be used in the
1652: construction we are ready to define the model.
1653:
1654: \subsection{The model}
1655:
1656: There are now several possible models that can be formulated, depending on
1657: which $\theta$-projector one uses. We shall proceed at the formal
1658: level, introducing factors of $\eta$ when necessary. Later the models
1659: we obtain will be placed on a more solid footing using the chain mail
1660: techniques.
1661:
1662: Let us consider a projector operator $\eta^2 \theta$.
1663: The basic field of the model
1664: is $\phi\in {\mathcal D}^{\otimes 3}$ that is real $\phi^*=\phi$, and
1665: one builds a projected field $\tilde{\phi}$ via:
1666: \be\label{theta-phi}
1667: &{}&\tilde{\phi}:=\eta^2 \theta\bullet \phi=\eta^2 \,\,
1668: \lower0.25in\hbox{\epsfig{figure=c-phi.eps, height=0.5in}}
1669: \ee
1670: In this graphical notation what is inserted on the ``meridian'' link
1671: and other 3 strands depends on the model. In all cases $\eta^2 \theta$
1672: is a projector. Using the graphical notation introduced, the
1673: action for the model is defined as follows:
1674: \be\label{action*}
1675: S[\phi] = \frac{\eta^2}{2} \,\,\, \lower0.25in\hbox{\epsfig{figure=c-kinetic.eps, height=0.5in}}
1676: + \frac{g}{4!} \eta^8 \,\,\, \lower1in\hbox{\epsfig{figure=c-inter.eps, height=2in}}
1677: \ee
1678:
1679: To describe Feynman amplitudes generated by these models we shall introduce a notion of
1680: the chain mail.
1681:
1682: \subsection{Roberts' chain mail}
1683:
1684: Using an idea of chain mail Roberts presented \cite{Roberts} a very convenient description of the
1685: Turaev-Viro invariant. Here we remind the reader this notion, and show that the Feynman
1686: amplitude generated by our model is just a chain mail evaluation.
1687:
1688: Let us consider the Feynman diagram perturbative expansion of the
1689: model \eqref{action*}. As the propagator of the model consists of
1690: 3 lines, each diagram is a collection of vertices (0-cells), edges
1691: (1-cells) and faces (2-cells) obtained by following each line
1692: till it closes. Each diagram $D$ is an abstract one, that is not embedded
1693: in any space. Note that the data of a diagram $D$ define a
1694: handlebody $H(D)$, which is just the blow up of the graph of $D$.
1695: In other words, to obtain $H(D)$ one takes solid balls (0-handles), one for every
1696: vertex of the diagram, and attaches them to each other by solid
1697: cylinders (1-handles), one cylinder for every edge. The handlebody $H(D)$ is not
1698: embedded in any space.
1699:
1700: Now draw the curves $\epsilon_i$ defining the 2-faces on the
1701: boundary $\partial H(D)$, and push them slightly inside of $H(D)$. Let us
1702: then add the meridian curves $\delta_i$ for all the 1-cells. The
1703: obtained collection of curves is called a {\it chain mail link}
1704: $C(D)\subset H(D)$ of $D$. We now attach an appropriate
1705: (model-dependent) operator from $\mathcal D$ to every component of the
1706: link $C(D)$, and evaluate the link to obtain a value $\Omega C(D) \in
1707: \CC$. This evaluation procedure is as follows. One views the
1708: chain mail as a rule for taking a product of $\theta$-projectors, one
1709: $\theta$ for every 1-cell. The projectors are $\bullet$-multiplied as specified
1710: by the 0-cells. Every closed loop corresponds to applying the Haar
1711: functional $h$. This evaluation rule will become more clear when we consider concrete examples
1712: below. We note that this evaluation is {\it not} an evaluation
1713: of the chain mail as embedded in some 3-manifold. It can be related to
1714: the more standard evaluation of the chain mail as embedded (as appears
1715: in Roberts' work \cite{Roberts}), see more remarks on this below. At
1716: the moment, however, there is no 3-manifold associated to any Feynman diagram.
1717:
1718: \begin{proposition} \label{prop} For the model \eqref{action*} the amplitude
1719: $\mathcal Z$ of a Feynman diagram $D$ is equal to $g^V \eta^{2E}
1720: \Omega C(D)$, where $V$ is the number of vertices and $E$ is the
1721: number of edges of $D$. Feynman amplitude of the model constructed
1722: using $\hat{1}$ is equal to $g^V \Omega C(D)$.
1723: \begin{proof} In the space of projected fields
1724: $\tilde{\phi}=\theta\bullet\phi$ the kinetic operator $\eta^2
1725: \theta$ is the identity operator. Its inverse is itself. Thus, the
1726: propagator of the model \eqref{action*} is
1727: again $\eta^2 \theta$, and the vertex is $g$ times the operator
1728: depicted graphically in \eqref{action*}. This set of Feynman rules
1729: makes the first statement evident. The second case is similar,
1730: except that no factors of $\eta$ appear.
1731: \end{proof}
1732: \end{proposition}
1733:
1734:
1735: \subsection{Feynman diagrams and 3-manifolds}
1736:
1737: Each Feynman diagram $D$ defines a 3D pseudo-manifold $PM(D)$. It is easiest to obtain
1738: $PM(D)$ by gluing together truncated tetrahedra an example of which is shown in Fig. \ref{fig:trun}.
1739: Thus, each vertex of $D$ naturally corresponds to a truncated tetrahedron. Each edge of $D$ defines a
1740: gluing of two truncated tetrahedra $T,T'$: a large face of $T$ is glued to a large face of $T'$. This way
1741: each (closed) Feynman diagram $D$ defines a 3D pseudo-manifold $PM(D)$ with boundary $\partial PM(D)$. The
1742: pseudo-manifold $PM(D)$ can be completed to a manifold $M(D)$ if the boundary $\partial PM(D)$ is a disjoint
1743: union of a number of spheres $S^2$. In this case $M(D)$ is formed by gluing to $PM(D)$ a number of 3-balls.
1744: See \cite{Roberto} for more details on this construction.
1745:
1746: Note, however that it may well happen for some of the Feynman diagrams that $\partial PM(D)$ contains
1747: boundary components other than spheres. In this case one needs extra structure to be added to the
1748: Feynman diagram to ``convert'' it into a 3-manifold. This extra structure is easiest to understand
1749: on the example of a toroidal boundary component in $\partial PM(D)$. In this case, to convert the
1750: 3-manifold $PM(D)$ with boundary into a closed 3-manifold one has to glue in a solid torus. However,
1751: there is no unique solid torus. A solid torus is specified by a closed curve on its boundary (torus) which
1752: is contractible inside the solid torus. Such a curve is in turn specified by two relatively prime
1753: numbers. This is exactly the extra data (for each toroidal boundary component)
1754: that are necessary to convert a Feynman diagram into
1755: a closed 3-manifold in cases when the manifold obtained by gluing in the two handles contains
1756: toroidal boundary components. When the boundary components are of higher genus, one needs
1757: even more data, which are a set of $3g-3$ curves on the boundary ($g$ is the genus of the boundary component
1758: in question) that are contractible inside the handlebody one glues in. These issues are delicate ones,
1759: and we would not like to divert the reader from the main theme by pursuing them any further.
1760: Let us just say that one either chooses (if possible) the symmetry properties of the model so that
1761: the boundary of $PM(D)$ is always a collection of spheres, or, if the former is not possible,
1762: adds some extra structure to the model that tells one how to complete the Feynman diagrams
1763: into 3-manifolds. We will assume that one of the two is done, and proceed without worrying
1764: about these issues any more.
1765:
1766: \begin{figure}\label{fig:trun}
1767: \begin{center}
1768: \epsfig{figure=trun-tet.eps, height=1.4in}
1769: \end{center}
1770: \caption{A truncated tetrahedron}
1771: \end{figure}
1772:
1773: Let us now explain a relation to the chain mail construction of the previous subsection is as follows. One obtains $PM(D)$
1774: by taking the handlebody $H(D)$ and gluing to it a number of 2-handles (solid cylinders). The
1775: gluing is done as specified by the $\epsilon_i$ curves. Once $PM(D)$
1776: is obtained, one glues a number of 3-handles to obtain $M(D)$. After this,
1777: using the ``extra rules'' discussed in the previous paragraph, one
1778: glues in a collection of 2-spheres or higher genus handlebodies, and completes $PM(D)$
1779: to a 3-manifold $M(D)$ without boundary. The original Feynman diagram $D$ is
1780: now embedded into $M(D)$. The evaluation that we described above did
1781: not need a representation of $D$ as embedded into some 3-manifold, and
1782: in that sense was independent of any embedding. However, once
1783: the diagram is embedded into $M(D)$, the evaluation that determines
1784: the Feynman amplitude can also be thought as the usual Reshetikhin-Turaev-Witten
1785: evaluation of $D$ in $M(D)$. The result of this evaluation does depend on
1786: the embedding of $D$ in $M(D)$. The amplitude one gets is thus that
1787: for a process described by the diagram $D$, happening in a particular
1788: 3-manifold $M(D)$. This should be contrasted with the usual situation
1789: in QFT, where the Feynman diagrams are not embedded in any space. The
1790: extra structure of faces and ordering of the faces that naturally follows
1791: from the group field theory expansion is exactly the structure that
1792: allows to think of $D$ as embedded in a 3-manifold. The 3-manifold
1793: that gets constructed from $D$ with its extra structure is what
1794: describes the gravitational part of the degrees of freedom of the model.
1795:
1796: We now note that different Feynman graphs can lead to one and the same topological
1797: manifold $M(D)$. There are certain moves that relate different $D$
1798: that correspond to the same manifold $M$. As stated by the following
1799: theorem, when the operator that is inserted on the strands satisfies
1800: the handleslide and killing properties, the chain mail evaluation is
1801: invariant under the moves and thus gives an invariant of $M$.
1802:
1803: \begin{theorem} (Roberts) Assume that the operator ${\mathcal O}$ inserted on links
1804: forming $C(D)$ satisfies the handleslide and killing properties. Assume
1805: that the trace of this operator is given by $\eta^{-2}$. The
1806: following multiple of the chain mail evaluation:
1807: \be\label{ch}
1808: CH(D):=\eta^{2n_0 + 2n_3} \Omega C(D),
1809: \ee
1810: where $n_0$ and $n_3$ are the numbers of 0 and 3-handles correspondingly,
1811: is an invariant of 3-manifolds in that $CH(D)=CH(D')$ when $M(D)\sim
1812: M(D')$ are 3-manifolds of the same topology.
1813: \begin{proof} We shall only present a sketch, as a detailed proof is
1814: given in \cite{Roberts}. Different Feynman diagrams that correspond
1815: to the same $M$ are different handle decompositions of $M$. A known
1816: theorem of topology states that different handle decompositions can
1817: be related by a sequence of births of $k, k-1$-handle pairs, and by
1818: handleslides. Invariance under handleslides is guaranteed by the the
1819: handleslide property of the operator inserted along the strands. To
1820: analyze the birth of 1-0 and 3-2 handle pairs one uses the
1821: handleslide property, and is left with a single unknot with
1822: ${\mathcal O}$ inserted and not linked with the rest of the chain
1823: mail. This gives a factor of $\eta^{-2}$ absorbed by the prefactor
1824: of $CH(D)$. A birth of the 2-1 handle pair is handled using the
1825: killing property.
1826: \end{proof}
1827: \end{theorem}
1828:
1829: \begin{remark} Note that if the operator inserted along the meridians
1830: and longitudes satisfies the handleslide and killing properties, and
1831: its trace is equal to unity, then a similar theorem holds, except
1832: that it is now the chain mail evaluation itself that is invariant.
1833: \end{remark}
1834:
1835: \begin{corollary}The model constructed using
1836: the identity operator $\delta_e(y)$ gives invariants of
1837: 3-manifolds.
1838: \end{corollary}
1839:
1840: \begin{remark} There is another model that gives manifold invariants, namely the model
1841: constructed using $\hat{1}$. This model, however, requires a somewhat different set of
1842: formal prefactors. We shall describe it in details below.
1843: \end{remark}
1844:
1845: \section{More general models}
1846: \label{pp-models}
1847:
1848: In the previous section we have seen how group field theory for the
1849: Drinfeld double leads to the notion of chain mail in that Feynman diagrams of GFT
1850: are computed as the chain mail evaluation. Chain mail arises because
1851: operators that are used in the construction of the GFT model are
1852: projectors. One can therefore multiply the $\theta$-projectors
1853: appropriately, and be left with a chain mail where there is just one
1854: meridian link per edge of a Feynman diagram $D$. Also importantly, the
1855: operator that is inserted in the strands is a projector, and after
1856: strands close combines to give a single operator in the longitude.
1857:
1858: One can consider a more general class of models not related to any
1859: GFT, but formulated directly in terms of the chain mail. Thus, one
1860: takes a chain mail that corresponds to a graph $D$, inserts one
1861: species of operators into the meridians, some other species into
1862: the longitudes, and evaluates the resulting link. The operators
1863: inserted do not have to be projectors anymore, and this is what makes such
1864: models different from the GFT ones. However, as we shall see, these more
1865: general models admit a physical interpretation and are of interest. We
1866: could have directly started from the notion of chain mail and
1867: formulated all models correspondingly. However, we believe that the
1868: GFT description we have given is essential in that it clear shows what
1869: is and what is not possible in the GFT framework. The GFT models
1870: described are also of importance in view of possible generalization to
1871: algebras other than $\D$. Thus, for instance, the analog of Boulatov
1872: model for the quantum group $\SU_q(2)$ has not been formulated.
1873: It is straightforward to do so using our algebraic formulation
1874: described above.
1875:
1876: \subsection{Formulation of the models}
1877:
1878: Let us define a set of models as follows.
1879: Consider a chain mail evaluation in which an operator $P=\delta_e(x) P(y)$ is
1880: inserted in all the longitudes. The function $P(y)$ that appears as part
1881: of this operator is not required to be a projector. Instead, we will just require $P(y)$ to be
1882: normalized so that $\int dy P(y)=1$. Let us consider the following model.
1883: \begin{definition} The amplitude of the model is
1884: obtained by inserting the identity $\delta_e(y)$ into the meridian
1885: links, and $\delta_e(x) P(y)$ into the longitudes.
1886: The amplitude is a (formal) multiple of the chain mail evaluation:
1887: \be\label{gen-model}
1888: CH'(D)=\eta^{2n_1} \Omega C(D).
1889: \ee
1890: Here $\Omega$ denotes the evaluation, and the factor of $\eta^2$ to the power of the
1891: number of 1-handles is introduced for reasons that will become clear below.
1892: \end{definition}
1893:
1894: Let us prove some properties of the objects that the both models are
1895: constructed from. First let us consider a
1896: number of strands enclosed by the $P$ operator. Such an object is
1897: given by:
1898: \be\label{p-1}
1899: P(x_1^{-1} \ldots x_n^{-1}) \prod_i \psi_i(x_i, y_i).
1900: \ee
1901: Importantly, there is no longer a handleslide property \ref{lemma-2} for
1902: objects \eqref{p-1}, as is easy to verify. However, the killing property still
1903: holds. Indeed, we take $\psi_n=\delta_e(y)$, and apply the Haar functional
1904: in the last channel. We get:
1905: \be
1906: \int dx_n P(x_1^{-1} \ldots x_n^{-1}) \prod_{i=1}^{n-1} \psi_i(
1907: x_i, y_i) \delta_e(e) = \\ \nonumber \eta^{-2} \int dg P(g) \, \prod_{i=1}^{n-1} \psi_i(x_i,y_i) =
1908: \eta^{-2} \, \prod_{i=1}^{n-1} \psi_i(x_i,y_i) ,
1909: \ee
1910: where we made a change of integration variable to arrive to the second expression.
1911:
1912: Using these facts, it is easy to prove the following assertion.
1913: \begin{proposition}\label{prop*} Amplitudes given by \eqref{gen-model} are invariant
1914: under 0-1 and 2-1 handle pair births/deaths, as well as under 1-handle slides.
1915: \begin{proof} Proof is same as that of Roberts' theorem of the previous
1916: section. The factor of $\eta^{2n_1}$ in \eqref{gen-model} is
1917: necessary to guarantee the 0-1 handle pair births/deaths
1918: invariance as well as the invariance under the 2-1 handle pair births/deaths. Note that in
1919: the case considered by Roberts, see (\ref{ch}) above, one needs two prefactors to guarantee
1920: the 0-1 and 2-3 moves. In our case we do not require the 2-3 moves, but have an additional
1921: factor of $\eta^{-2}$ appearing in the 1-2 moves. This makes the correct prefactor
1922: to be $\eta^2$ to the power of $n_1$ not $n_0$ as in case of (\ref{ch}).
1923: \end{proof}
1924: \end{proposition}
1925:
1926:
1927: \subsection{A model giving 3-manifold invariants}
1928:
1929: Let us note that the case $P(y)=1$ is special, because in this case there
1930: is the 2-handleslide property. In this case the chain mail evaluation gives 3-manifold
1931: invariants. We have the following (almost trivial) corollary.
1932: \begin{corollary} Evaluate the chain mail by inserting the identity
1933: operator $\delta_e(y)$ in all meridian links, and the operator $\delta_e(x)$
1934: into all longitudes. The quantity \eqref{gen-model} is a topological
1935: invariant in that it only depends on the topology of $M(D)$. Moreover,
1936: its value is one for any manifold $M(D)$.
1937: \end{corollary}
1938:
1939: \subsection{Interpretation of 2-handleslide invariance absence}
1940:
1941: The example of the previous subsection shows that there are models
1942: that are interesting but not in the class of GFT models considered in
1943: the previous section. It is not hard to give a more flexible
1944: definition of GFT that would cover more examples. We shall not attempt
1945: this however, concentrating instead on the chain mail definition from
1946: now on.
1947:
1948: General models defined via chain mail do not produce
1949: 3-manifold invariants because there is no handleslide property for the
1950: operator inserted in the longitudes. Still, as we shall presently see,
1951: from the point of view of point particle theory these models are quite
1952: natural. To see this, let us analyze what
1953: happens if there is no handleslide property. Meridian links correspond to 1-handles,
1954: and the longitudes correspond to 2-handles. There is
1955: the handleslide property for the identity operator, see lemma
1956: \ref{lemma-2}. Thus, the chain mail evaluation $\Omega CH(D)$ is
1957: invariant under the 1-0 handle pair births and deaths. The killing
1958: property guarantees an invariance under the 2-1 pair births and
1959: deaths. There is also invariance under the handleslides along
1960: 1-handles. Thus, non-invariance comes only from two sources: there is
1961: no invariance under 2-handle slides, and there is no 3-2 handle pair
1962: birth-death invariance. Now recall that each $D$ is the dual skeleton
1963: of some triangulated 3-manifold $M(D)$. As we have discussed in the
1964: introduction, we would like to interpret the triangulation as a
1965: Feynman diagram as well. Edges of this Feynman diagram are in
1966: one-to-one correspondence with 2-handles, and vertices with 3-handles
1967: of $M(D)$. Absence of invariance under moves involving 2-handles just
1968: means that the amplitude we get is not invariant under changes in the
1969: Feynman diagram. The absence of invariance under 3-2 pair birth-death
1970: means that the Feynman amplitude is not invariant under a vertex being
1971: replaced by two vertices connected by a new edge. Similarly, the
1972: absence of the 2-handleslide property means that one cannot move edges
1973: through vertices along other edges. All these moves change the Feynman diagram, and
1974: the fact that there is no invariance is very natural from the point of
1975: view of point particle field theory. Thus, in this case the model
1976: leads not to 3-manifold invariants, but to Feynman amplitudes of the
1977: dual point particle field theory. To get most general point particle
1978: models we will have to work with theories of this class.
1979:
1980: \section{Evaluation and interpretation of the models}
1981: \label{sec:pp}
1982:
1983: We have considered a set of models. Some of them were shown to give
1984: invariants of 3-manifolds, some other only give amplitudes for
1985: dual Feynman graphs. To analyze the structure of the amplitudes
1986: arising, and to give it a geometrical and
1987: point particle interpretation, let us use the kernel representation.
1988:
1989: \subsection{Kernel representation}
1990:
1991: Recall that in the kernel representation one
1992: associates an operator $f(x,y)\in {\mathcal D}$
1993: its kernel defined as $(\kappa\otimes{\rm id})\Delta^* f$. A simple computation gives:
1994: \be\label{k-1}
1995: f(x,y|\tilde{x},\tilde{y})=f(y^{-1} x^{-1} y, y^{-1} \tilde{y}) \delta_e(x \tilde{x}).
1996: \ee
1997: The kernels should be multiplied using the $\star$-product and the $h^*$ Haar functional should be taken.
1998: In the case of algebra of functions on the group this operation is just a point-wise multiplication
1999: of the propagators with an integral taken. In the case of the Drinfeld double the structure is more involved.
2000: To display it, we note that:
2001: \be
2002: h^*(f_1\star f_2) = \int dxdy\,\, f_1(x,y)f_2(x^{-1},y).
2003: \ee
2004: Thus, in order to compose kernels of two operators, one should take one kernel and multiply it
2005: by the other kernel with $x$ replaced by $x^{-1}$, and then integrate over $x,y$. Equivalently, one
2006: can replace $x$ by $x^{-1}$ in \eqref{k-1}. It is this quantity that
2007: we shall refer to as the operator kernel:
2008: \be\label{k-2}
2009: K_f(x,y|\tilde{x},\tilde{y}):=f(y^{-1} x y, y^{-1} \tilde{y}) \delta_x(\tilde{x}),
2010: \ee
2011: One can now multiply the kernels $K_f$
2012: in the usual way, and integrate over the gluing variables. Thus, it is easy to check that:
2013: \be
2014: \int dpdq\,\, K_{f_1}(x,y|p,q) K_{f_2}(p,q|\tilde{x},\tilde{y})=K_{f_1\bullet f_2}(x,y|\tilde{x},\tilde{y}).
2015: \ee
2016: However, the kernel formalism leads to a singularity for the Haar
2017: functional $h$. Indeed, we have:
2018: \be
2019: \int dx dy\,\, K_f(x,y|x,y) = \delta_e(e) h(f) = \eta^{-2} h(f).
2020: \ee
2021: Thus, in order to reproduce the correct Haar functional the trace of
2022: the kernel should be renormalized by the already familiar quantity
2023: $\eta^2$ \eqref{eta}.
2024:
2025: \subsection{Group field theory Feynman diagram analysis}
2026:
2027: Let us first apply the kernel technique to the model that was
2028: constructed from the identity operator $\delta_e(y)$ and that gives
2029: 3-manifold invariants. In this subsection we group quantities
2030: according to GFT Feynman graphs $D$. In the next subsection we perform
2031: a ``dual'' analysis, in which a point particle interpretation is more
2032: clear.
2033:
2034: It is clear that the amplitudes can be computed as follows: one should
2035: take the kernel for each vertex of $D$, and compose them together as
2036: specified by the diagram. This is due to the fact that the propagator
2037: of the model is proportional to $\theta$, and the vertex in
2038: \eqref{action*} is invariant under multiplication by $\theta$ from
2039: any of the 4-possible directions, see \eqref{action*}. Let us analyze
2040: this vertex kernel and give its geometrical
2041: interpretation.
2042:
2043: First we need an expression for the $\theta$-projector
2044: kernel. Applying \eqref{k-2} to \eqref{theta*} we get:
2045: \be\label{k-3}
2046: K_\theta(x_1,y_1;x_2,y_2,x_3,y_3|\tilde{x}_1,\tilde{y}_1;\tilde{x}_2,\tilde{y}_2;\tilde{x}_3,\tilde{y}_3)=
2047: \delta_{x_1}(\tilde{x}_1) \delta_{x_2}(\tilde{x}_2)
2048: \delta_{x_3}(\tilde{x}_3) \\ \nonumber
2049: \delta_e(y_1^{-1} x_1 y_1 y_2^{-1} x_2 y_2 y_3^{-1} x_3 y_3)
2050: \theta(y_1^{-1} \tilde{y}_1,y_2^{-1} \tilde{y}_2,y_3^{-1}
2051: \tilde{y}_3).
2052: \ee
2053: Let us now find the vertex kernel. As in section \ref{sec:Boul}, we
2054: shall enumerate 4 channels of the vertex by indices
2055: $i,j=1,\ldots,4$. The vertex kernel is then a function of variables
2056: $x_{ij}, y_{ij}$, where $x_{ij}\not=x_{ji}, y_{ij}\not=y_{ji}$. The
2057: kernel is obtained by taking a composition of 4 kernels
2058: \eqref{k-3}. We have:
2059: \be
2060: V(x_{ij},y_{ij}) = \int \prod_{i\not=j=1}^4 d\tilde{x}_{ij}
2061: d\tilde{y}_{ij} K_\theta(x_{ij},y_{ij}|\tilde{x}_{ij},\tilde{y}_{ij}),
2062: \ee
2063: where $\tilde{x}_{ij}=\tilde{x}_{ji},
2064: \tilde{y}_{ij}=\tilde{y}_{ji}$. It is straightforward to do the
2065: $\tilde{x}$ integrations. Each kernel \eqref{k-3} contains a $\delta_x(\tilde{x})$
2066: factor, which makes the result proportional to
2067: $\delta_{x_{ij}}(x_{ji})$. Thus, the vertex kernel is actually a
2068: function of 6 variables $x_{ij}=x_{ji}$:
2069: \be\label{k-4}
2070: V(x_{ij},y_{ij}) = \prod_{(i,j,k,l)} \delta_e( y^{-1}_{ij}
2071: x_{ij} y_{ij} y^{-1}_{ik} x_{ik} y_{ik} y^{-1}_{il} x_{il} y_{il}) \\
2072: \nonumber
2073: \int d\tilde{y}_{ij} \, \theta(y^{-1}_{ij} \tilde{y}_{ij},y^{-1}_{ik}
2074: \tilde{y}_{ik}, y^{-1}_{il} \tilde{y}_{il}).
2075: \ee
2076: Here $\tilde{y}_{ij}=\tilde{y}_{ji}$, so the integration is taken over
2077: 6 variables. We have also introduced a notation: $(i,j,k,l)$ for a
2078: quadruple of integers: $i\not=j\not=k\not=l$. The last $\tilde{y}$
2079: integral in \eqref{k-4} can be taken at the
2080: expense of introducing another 4 variables $g_i$. Indeed, let us use
2081: the formula \eqref{theta} for each of the 4 $\theta$-functionals
2082: in \eqref{k-4}. We get:
2083: \be\label{vertex-kern}
2084: V(x_{ij},y_{ij}) = \!\!\!\!\prod_{(i,j,k,l)} \!\!\delta_e( y^{-1}_{ij}
2085: x_{ij} y_{ij} y^{-1}_{ik} x_{ik} y_{ik} y^{-1}_{il} x_{il} y_{il})
2086: \int \prod_{i<j} dg_i \,\, \delta_e(y_{ij} g_i g_j^{-1} y^{-1}_{ji}).
2087: \ee
2088: We note that the last term in this expression is just the vertex
2089: kernel of the Boulatov model. Thus, the modification of the vertex as
2090: compared to the Boulatov model case is in an appearance of the 4
2091: additional $\delta$-function terms, and in the fact that the kernel
2092: depends on 6 variables $x_{ij}$.
2093:
2094: One can re-interpret this structure
2095: by saying that the only modification as compared to the Boulatov model
2096: case is that the edge contribution changed: there is now an extra
2097: $\delta$-function for each edge of $D$. Let us change the notation
2098: to display the structure arising more clearly. Let us denote the
2099: vertices, edges and faces of $D$ by tilded letters: $\tilde{v},
2100: \tilde{e}, \tilde{f}$ correspondingly. We leave the untilded letters
2101: for elements of the triangulation $\T$ to which $D$ is dual.
2102: Let us now denote the
2103: $y$-variables as $y_{\tilde{e},\tilde{f}}$, where $\tilde{e}$ is an
2104: edge, and $\tilde{f}$ is a face of
2105: $D$ (each edge $\tilde{e}$ is shared by exactly 3 faces). Let us denote the
2106: $x_{ij}$ variables by $x_{\tilde{f}}$. Indeed, because there is a
2107: $\delta$-function imposing $x_{ij}=x_{ji}$, all $x$-variables
2108: belonging to edges around a face are equal, so there is just one
2109: $x$-variable per face. Thus, the edge factor becomes:
2110: \be\label{edge}
2111: \delta_e(\prod_{\tilde{f}} y^{-1}_{\tilde{e},\tilde{f}} x_{\tilde{f}}^{-1} y_{\tilde{e},\tilde{f}})
2112: \ee
2113: for each edge $\tilde{e}$ of $D$. The vertex, on the other hand, is given by the
2114: same expression as in the case of Boulatov model:
2115: \be\label{vertex-boul}
2116: \int \prod_{\tilde{e},\tilde{e}'\in \tilde{v}} dg_{\tilde{v},\tilde{e}} \,\,
2117: \delta_e(y_{\tilde{e},\tilde{f}} g_{\tilde{v},\tilde{e}}
2118: g_{\tilde{v},\tilde{e}'}^{-1} y^{-1}_{\tilde{e}',\tilde{f}}),
2119: \ee
2120: where for each pair $\tilde{e},\tilde{e}'$ of edges that emanate from vertex $\tilde{v}$ $\tilde{f}$
2121: is the face that shares both $\tilde{e},\tilde{e}'$. Note that the
2122: $g$-variables are different at different vertices
2123: $g_{\tilde{v},\tilde{e}}\not=g_{\tilde{v}',\tilde{e}}$. To obtain the amplitude one
2124: multiplies the edge and vertex kernels, and integrates over the
2125: $y_{\tilde{e},\tilde{f}}, x_{\tilde{f}}$ variables.
2126:
2127: Already at this stage an interesting geometric interpretation of all
2128: the variables is possible. Indeed, we have already encountered the
2129: $g_{\tilde{v},\tilde{e}}$ variables before. Recall that each Feynman diagram $D$ is dual to
2130: a triangulated 3-manifold, and therefore each vertex $\tilde{v}$ is dual to a
2131: tetrahedron $t$. Thus, all the variables in \eqref{vertex-boul} are those
2132: of a single tetrahedron. Let us denote the triangulation to which $D$
2133: is the dual graph by $\T$. We denote the elements of $\T$ by untilded
2134: letters. Thus, $v,e,f,t$ are vertices,
2135: edges, faces and tetrahedra of $\T$ correspondingly. We see that
2136: $g_{\tilde{v},\tilde{e}}$ are variables of a tetrahedron $t$ dual to
2137: $\tilde{v}$ and correspond to
2138: faces $f$ of $\T$. Each $g_{\tilde{v},\tilde{e}}$ has the interpretation of
2139: the holonomy $g^t_f$ of the connection
2140: across the face $f$. There is a similar variable in the neighboring
2141: tetrahedron $t'$ and the total holonomy across the face $f$ is given by the
2142: product: $g_f=(g^t_f)^{-1} g^{t'}_f$. As we have already discussed in the
2143: introduction, the product $(g^t_f)^{-1} g^t_{f'}$ describes a
2144: rotation of one face $f$ into the other $f'$ and thus carries
2145: information about the dihedral angle between these faces.
2146:
2147: \begin{figure}\label{fig:2tet}
2148: \begin{center}
2149: \epsfig{figure=2-tetra.eps, height=1.4in}
2150: \end{center}
2151: \caption{A compound of a tetrahedron $t$ with a truncated tetrahedron $\bar{t}$.}
2152: \end{figure}
2153:
2154: The variables $y_{\tilde{e},\tilde{f}}$ are more interesting and this is the first time
2155: that we have encountered them. Using the duality, we can also write
2156: them as $y_{e,f}$. There is 12 of them for each
2157: tetrahedron. It is natural to interpret
2158: them in terms of a certain other truncated tetrahedron. Indeed, let us introduce
2159: the dual tetrahedron, and cut off its vertices by surfaces
2160: parallel to the faces of $t$ to obtain a truncated tetrahedron
2161: $\bar{t}$. A compound object made of $t$ and
2162: $\bar{t}$ is shown in Fig. \ref{fig:2tet}. The truncated tetrahedron $\bar{t}$
2163: has 12 vertices, and we shall interpret 12 variables $y_{e,f}$ as
2164: describing ``positions'' of these vertices. The meaning of each
2165: $\delta$-function in \eqref{vertex-boul} is then as follows: it imposes the constraint that
2166: the rotation of a vertex $y_{e,f}$ into
2167: $y_{e,f'}$ around the edge $e$ is the same as rotation of the face $f$
2168: into $f'$ as described by $g$-variables.
2169:
2170: Thus, we
2171: only have the variables $x_{\tilde{f}}$ uninterpreted. By duality,
2172: they are variables $x_e$ that have to do with the edges of $t$. Tetrahedron
2173: $t$ is truncated, and we interpret $x_e$ as describing rotations of the small
2174: faces of the truncated tetrahedron $t$ into one another. Thus,
2175: interpretation of all of the variables in \eqref{vertex-kern} is in
2176: terms of geometrical quantities associated with two truncated
2177: tetrahedra $t, \bar{t}$ inscribed into one another.
2178:
2179: For the convenience of the reader, let us re-write the above edge and
2180: vertex factors using the triangulation $\T$ notations. We have:
2181: \be\label{edge'}
2182: \delta_e(\prod_e y^{-1}_{e,f} x_e^{-1} y_{e,f})
2183: \ee
2184: for each face $f$ of $\T$. Here the product in the argument of the
2185: $\delta$-function is taken over 3 edges that form the boundary of face
2186: $f$. The tetrahedron factor is given by:
2187: \be\label{vertex-boul'}
2188: \int \prod_f dg^t_f \prod_{f,f'\in t} \,\,
2189: \delta_e(y_{e,f} g^t_f
2190: (g^t_{f'})^{-1} y^{-1}_{e,f'}).
2191: \ee
2192:
2193: The interpretation that we would like to propose for the total
2194: tetrahedron amplitude
2195: \eqref{vertex-kern} is that of a wave function of a pair of dual to
2196: each other truncated tetrahedra. To obtain an amplitude for a manifold
2197: one multiplies these wave functions and integrates over the gluing
2198: variables $x,y$. We would like to note that there is a similarity in
2199: the structure of the vertex found and the expression for the
2200: $6j$-symbol obtained in \cite{Holography}. In both cases the quantity
2201: in question has to do with a truncated tetrahedron, and is constructed
2202: using propagators for long edges (the second set of $\delta$-functions
2203: in \eqref{vertex-kern}), and a set of factors for the small faces (the
2204: first set of $\delta$-functions in \eqref{vertex-kern}). Thus, the
2205: structure we have found is probably pertinent to a very general class
2206: of models.
2207:
2208: If not for the edge factors \eqref{edge'}, one could easily integrate over the $y$ variables with the result
2209: being a product of $\delta$-functions, one for every edge $e$ of $\T$, and requiring that the holonomy
2210: around $e$ is trivial. With factors \eqref{edge'} the integration is not straightforward. The total
2211: amplitude \eqref{vertex-kern} is one where tetrahedron constraints discussed in the introduction are taken
2212: into account.
2213:
2214: \subsection{Dual graph analysis}
2215:
2216: Above we have represented the chain mail evaluation $\Omega CH(D)$ as
2217: a composition of vertex kernels \eqref{vertex-kern}, or the edge
2218: \eqref{edge} and vertex \eqref{vertex-boul} factors. Edges and
2219: vertices are those of $D$. We would now like to group the factors
2220: according not to $D$ but to a triangulation $\T$ whose dual graph is
2221: $D$.
2222:
2223: Thus, Feynman diagrams $D$ have an
2224: interpretation of the dual skeleton of some triangulated
2225: 3-manifold. Let us denote this triangulation by $\T$. Let us consider
2226: vertices $v$ of this
2227: triangulation. The edges $e$ of $\T$ are formed by parts of the meridian
2228: links of $CH(D)$. The number of strands forming each edge of $\T$
2229: is not fixed, and is equal to the number of faces of $\T$ sharing
2230: this edge. Strands forming the edges of Feynman diagram $D$, and whose
2231: closure defines the faces of $D$, now enclose the strands forming the
2232: edges of $\T$. Thus, we have a very similar structure to that
2233: considered above, except that it is a dual one, and that the number of
2234: strands enclosed by a link is now not fixed. To evaluate $CH(D)$ we
2235: have to insert the identity operator in each strand. Thus, we insert
2236: in each edge $e$ of $\T$ the
2237: following operator:
2238: \be
2239: \delta_e(x_1^{-1}\ldots x_n^{-1}) \int dx\, \prod_i
2240: \delta_e(x^{-1} y_i).
2241: \ee
2242: As before, it is convenient to introduce the kernel \eqref{k-2}:
2243: \be\label{int-1}
2244: \delta_e(y_1^{-1} x_1^{-1} y_1 \ldots y_n^{-1}
2245: x_n^{-1}y_n) \int dx\, \prod_i
2246: \delta_x(y_i^{-1} \tilde{y}_i) \delta_{x_i}(\tilde{x}_i).
2247: \ee
2248: As in the previous subsection we can take a composition of such
2249: kernels around every vertex of $\T$ by integrating over the
2250: $\tilde{x},\tilde{y}$ variables. Taking into
2251: account the $\delta_{x_i}(\tilde{x}_i)$ functions, we see that there
2252: remains a single variable $x_f$ for every face. The $y$-variables can be
2253: denoted by $y_{e,f}$. Indeed, there is one such variable for each
2254: strand forming each edge of $\T$, or, equivalently,
2255: for each face $f$ sharing edge $e$. This vertex kernel is found to be:
2256: \be\label{v-kern}
2257: \prod_e \delta_e(\prod_{f} y_{e,f}^{-1} x_f^{-1} y_{e,f})
2258: \prod_{e,e'} \delta_e(y_{e,f} x_{v,e} x_{v,e'}^{-1} y_{e',f}^{-1}).
2259: \ee
2260: Here the second product is over pairs of edges that emanate from $v$,
2261: and $f$ in the argument of the second set of $\delta$-functions is the face
2262: shared by both $e,e'$. Equivalently, one can say that second first
2263: product is over the faces that touch $v$.
2264: To obtain the chain mail evaluation one should multiply the
2265: kernels \eqref{v-kern} and integrate over all the variables $x_{v,e}, x_f,
2266: y_{e,f}$. Note
2267: that the vertex \eqref{v-kern} is essentially the same as
2268: \eqref{vertex-kern}, except that the dual quantities are used.
2269:
2270: Let us discuss a geometrical interpretation of the quantities involved
2271: in \eqref{v-kern}. The quantity $x_f$ is interpreted as the
2272: holonomy across the face $f$. The quantity $y_{e,f}$ encodes the
2273: dihedral angles of a tetrahedron. Namely, $y_{e,f} y_{e,f'}^{-1}$
2274: describes the rotation of face $f$ into $f'$ around edge $e$. We see
2275: that arguments of the first set of $\delta$-functions in \eqref{v-kern} are
2276: just the product of holonomies across the faces that share $e$ with the group
2277: elements that describe rotations of $f$ to the next face $f'$. Thus,
2278: the argument of the $\delta$-function for edge $e$ is the total
2279: rotation that one gets by going around
2280: $e$. This means that there is no deficit angle allowed! Thus, in spite
2281: of the fact that the model was constructed using $\D$, not just
2282: $\SU(2)$, the model does not
2283: seem to describe point particles. We shall return to this conundrum
2284: below.
2285:
2286: For now let us note that an interpretation of the second set of $\delta$-functions in
2287: \eqref{v-kern} is also possible. Indeed, we see that the quantity $x_{v,e}
2288: x_{v,e'}^{-1}$ describes the rotation of edge $e$ into $e'$, with the
2289: vertex $v$ as the center of rotation. Each $\delta$-function
2290: guarantees that $y_{e,f}^{-1} y_{e',f}$ is the same as the rotation
2291: that sends $e$ to $e'$.
2292:
2293: \subsection{General models and their point particle interpretation}
2294:
2295: As we have just seen, the topological GFT model constructed from the
2296: identity operator $\delta_e(y)$ on the Drinfeld double does not
2297: describe point particles. Earlier we have introduced a more general
2298: set of models in which a function $\delta_e(x) P(y)$ is inserted into the
2299: merideans. Now we would like to show that these models do describe
2300: point particles and $P(y)$ is the particle's propagator in the
2301: momentum space.
2302:
2303: We have seen in the previous subsection, the point particle
2304: interpretation, if any, is most visible in a description in which one
2305: groups all the projectors around vertices of $\T$. Let us give this
2306: description. As before, each
2307: edge of $\T$ is formed by a number of strands with the identity
2308: operator inserted. However, the operator on the link that encloses the
2309: edge is now $P$. Therefore, each edge of $\T$ corresponds to the
2310: following operator:
2311: \be
2312: P(x_1^{-1}\ldots x_n^{-1}) \prod_i \delta_e(y_i).
2313: \ee
2314: These operators now have to be multiplied (using the $\bullet$-product) as specified by the fat
2315: structure, and the Haar functional has to be applied in each strand to get the amplitude. It is
2316: easiest to understand what the result is by noticing that each of these edge operators is
2317: actually an element of the ideal $({\mathcal A}^*)^{\otimes n}$. Indeed, functions of the form
2318: $f(x) \delta_e(y)$ form an ideal in $D(G)$, isomorphic to the algebra ${\mathcal A}^*$ with
2319: point-wise multiplication. Therefore, the result of the $\bullet$-product of all
2320: these operators is the point-wise product of the functions $P(x_1^{-1}\ldots x_n^{-1})$
2321: times the product of the $\delta$-functions $\delta_e(y)$. It is now clear what the evaluation
2322: is. Let us first discuss what happens with the $y$-dependence. Applying the Haar functional
2323: $h$ in each strand we get an infinity $\delta_e(e)$. These infinities are cancelled
2324: by the prefactor $\eta^{2n_1}$ that we introduced in (\ref{gen-model}). One could,
2325: actually, avoid this infinities altogether by a slight modification of the evaluation procedure.
2326: Thus, since the result in each strand is always in ${\mathcal A}^*$ ideal of $D(G)$, let us
2327: instead of applying the Haar functional of $D(G)$ apply the one of ${\mathcal A}^*$.
2328: The result of this evaluation procedure is finite (for sufficiently nice propagators for which
2329: the group integrals would converge).
2330:
2331: The result of this evaluation procedure is given by
2332: \be
2333: \int \prod_f dx_f \prod_e P(x_{f_1}^{-1} \ldots x_{f_n}^{-1}).
2334: \ee
2335: Here the first product is over the faces of the diagram, and $x_f$ is a group variable
2336: that gets associated to each face. The second product is over the edges of the diagram,
2337: and the product of $x$-variables in the argument of the propagator includes these variables
2338: for the faces $f_1, \ldots, f_n$ shared by that edge. It is clear now that the function
2339: $P(y)$ plays the role of the particle propagator, for the above answer for the amplitude is as
2340: expected, see e.g. \cite{Barrett,Freidel}.
2341:
2342: Thus, we have arrived at the expression we started from in the
2343: Introduction. Indeed, the amplitude is just the Feynman amplitude in
2344: the momentum representation, with the propagator evaluated at
2345: the product of face variables around an edge. The momentum
2346: conservation constraints are solved by introducing the face
2347: variables. The tetrahedron constraints that we discussed in the
2348: Introduction can be omitted because the integrand is invariant under
2349: the action of the gauge transformations at vertices of the dual
2350: triangulation, as can be easily checked. Moreover, as the proposition
2351: \ref{prop*} states, the Feynman amplitude is to a large extent
2352: independent of the extra ``fat'' structure that is added to the
2353: diagram in order to define it.
2354:
2355: There are two interesting ``limiting cases'' to consider. The case $P(y)=\delta_e(y)$
2356: corresponds to having no particles at all. This gives the topological Ponzano-Regge model,
2357: if one introduces an additional prefactor $\eta^{2n_3}$ that is required to guarantee
2358: the invariance under the 2-3 moves. The other case is $P(y)=1$, which gives another
2359: (trivial) topological model, the one that gives one for every manifold as the amplitude.
2360: However, one can decompose the function $P(y)=1$ into $\delta$-functions on
2361: conjugacy classes as in (\ref{ident-1}). This shows that even this trivial model
2362: can be thought of as containing particles of all possible masses in it, but
2363: becoming trivial after the integral over the particles' mass is taken.
2364:
2365: \section{Discussion}
2366:
2367: Let us recap the main points of this work. We have considered the group
2368: field theory for the Drinfeld double and constructed an exact analog
2369: of the Boulatov model. The Feynman expansion of the model leads to
2370: triangulated 3-manifolds weighted with a manifold invariant. The
2371: invariant is the ``Ponzano-Regge'' model for the Drinfeld
2372: double. By reducing this Feynman amplitude to the chain mail
2373: evaluation we proved that this Drinfeld double Ponzano-Regge amplitude
2374: is indeed an invariant of 3-manifolds (when dressed by an appropriate
2375: power of the formal quantity $\eta^2$). We have also found that this
2376: model does not describe point particles. Indeed, one of the factors
2377: appearing in the amplitude for this model is a $\delta$-function for
2378: the holonomy around each triangulation edge. This is a bit
2379: disappointing, because our original motivation for considering the
2380: Drinfeld double GFT was precisely to obtain a model containing point
2381: particles. Thus, in a certain sense, our original motivation has
2382: failed.
2383:
2384: Nevertheless, motivated by the GFT considerations, and especially by the chain mail
2385: construction, we have introduced a more general set of models that are
2386: defined using not GFT but directly in terms of the chain mail. We have considered only one
2387: such model, but others are possible. The data that goes into
2388: this model is an assignment of a propagator (function on the
2389: $\SU(2)$ group manifold) to each edge of the Feynman diagram. This
2390: propagator can be chosen different for different edges, as long as it
2391: is appropriately normalized. The resulting Feynman amplitude is invariant under moves
2392: that do not change the Feynman graph and the 3-manifold topology. This model
2393: describes point particles. In fact, it is exactly the
2394: naive model envisaged already in the Introduction. Our results on its
2395: invariance under the topological moves show that this model is
2396: well-defined and to a large extent independent on the ``fat''
2397: structure that was introduced to define it. The model
2398: models become topological if one choses $P(y)=1$. In this case it seems trivial
2399: as the amplitude it gives to any manifold is one. At the same time,
2400: even in this case the model does describe point particles. One
2401: just has to interpret the operator $\hat{1}$ as the integral over
2402: particle's mass, or using physics terminology, the integral over the
2403: internal momenta of particles.
2404:
2405: The model we have considered (the one constructed with a general
2406: propagator $P$ inserted into the longitudes) has an interpretation as
2407: describing pure 3d gravity coupled to
2408: point particles. As we have discussed in the introduction, there is
2409: another possible interpretation. Namely, one can view this model
2410: as describing point particles only, but with the backreaction on the
2411: metric taken into account. We have not discussed what
2412: would be the corresponding theory in the position space. Let us only
2413: note that non-commutativity of momentum implies non-commutativity in the position
2414: space as well. It would be of interest to formulate
2415: point particle theories described here in the position space, and understand what
2416: kind of action principle is responsible for them.
2417:
2418: Now that one has a model for 3d quantum gravity with point particles, various
2419: quantities of physical interest can be computed. However, certain
2420: additional developments are necessary in order to achieve this. First it is
2421: important to be able to compute not only closed but also open Feynman graph
2422: amplitudes. This would give correlation
2423: functions. Second and more important, one should find an analog of LSZ
2424: reduction formula of field theory that expresses scattering amplitudes
2425: in terms of field correlation functions. We shall not attempt this in the present paper.
2426:
2427: A remark is in order about the topological invariant defined by our particle model.
2428: As we have already mentioned, this is ${\mathcal Z}=1$ for any manifold. However, the
2429: amplitude for each Feynman diagram (with a fixed set of conjugacy classes
2430: $\theta_e$ on the edges) is non trivial. Moreover, an amplitude for
2431: an open Feynman graph, which describes particle scattering and which
2432: we have not considered in this paper is also not trivial, even when
2433: one integrates over the internal momenta. This shows that, in spite of
2434: seeming triviality, this model is physically interesting.
2435:
2436: Let us now return to the topologically invariant model constructed from the
2437: identity operator $\delta_e(y)$. Several comments are in order. First,
2438: it would be interesting to understand
2439: how different this model is from that of Ponzano-Regge.
2440: The original Ponzano-Regge model imposes
2441: the constraints that the holonomy either around edges or dual edges is trivial. The model constructed
2442: in this paper imposes both sets of constraints at the same time. For
2443: the case of the double of a quantum $\SU_q(2)$, which makes the model
2444: well-defined (by rendering $\eta^{-2}$ finite), one expects the
2445: Drinfeld double ``Ponzano-Regge'' model to be equal to the usual
2446: Ponzano-Regge model amplitude squared. It would be interesting to
2447: understand the interpretation of the classical group Drinfeld double
2448: case better.
2449:
2450: Another remark is that, as is clear from our construction, the gauge symmetry
2451: $\delta\phi=\psi, \theta\bullet\psi=0$ of the group
2452: field theory is extremely important. From the chain mail description
2453: it is clear that topological invariance of the model is intimately
2454: related to the fact that $\theta$-projector
2455: is used as the propagator of the model. On the other hand, we know
2456: that topological invariance of 3d gravity is due to large amount of
2457: gauge symmetry present in it. More precisely, diffeomorphisms are so
2458: strong in 3d, that they render the theory topological. All this
2459: strongly suggests that there is a link between the group field theory
2460: and gravity gauge symmetries. In the context of 2d gravity matrix
2461: models the relation between gravity and a matrix model is an instance
2462: of open-closed string duality, in which one and the same Feynman
2463: amplitude has two different interpretations. We expect that the
2464: relation between group field theory and 3d gravity explored here is of
2465: a similar nature, and that the group field theory gauge symmetry
2466: becomes diffeomorphism symmetry on the gravity side. It would be of
2467: considerable interest to explore all this in more detail.
2468:
2469: It is clear that the GFT construction that we have presented is very
2470: general, and can be applied to algebras other than the Drinfeld
2471: double. The Drinfeld double of $\SU(2)$ is described \cite{Koorn} as the twisted product of the
2472: algebra of functions on $\SU(2)$ with its dual, which is also algebra
2473: of functions on $\SU(2)$. The Drinfeld twisted product construction
2474: can be applied to groups other than $\SU(2)$. Another important for
2475: physical applications example is given by the quantum Lorentz group,
2476: which is just the twisted product of the quantum $\SU(2)$ with $q$
2477: real and its Pontryagin dual, see \cite{Woronowicz}. Our construction
2478: can be applied to this case. We expect that in this case there is not
2479: one, but several possible models that lead to 3-manifold invariants,
2480: depending on which ``identity'' operator is used. These models should
2481: have the physical interpretation as describing point particles coupled
2482: to quantum gravity with either positive cosmological constant and
2483: Lorentzian signature or with negative cosmological constant and
2484: Euclidean signature. It would be of great interest to find all these
2485: models.
2486:
2487:
2488:
2489:
2490: \section*{Acknowledgments}
2491:
2492: I would like to thank my colleagues John Barrett and Jorma Louko for
2493: stimulating discussions and support. I am grateful to G. 't Hooft
2494: for a discussion about point particles in 2+1 gravity, and to
2495: L. Freidel for many important discussions about group field theory.
2496: The author is supported by an EPSRC Advanced Fellowship.
2497:
2498:
2499: \section*{Appendix: Some useful formulas}
2500:
2501: In the main text we often use the integral of two and three matrix elements. For two matrix elements we have:
2502: \be\label{A-2-1}
2503: \int dg\,\, \overline{T^j_{mn}(g)} T^{j'}_{m'n'}(g) = \frac{\delta_{jj'}}{{\rm dim}_j} \delta_{mm'} \delta_{nn'}.
2504: \ee
2505: In graphical notation
2506: this takes the following form:
2507: \be\label{A-2}
2508: \int dg \,\,\, \lower0.25in\hbox{\epsfig{figure=2g.eps, height=0.5in}} = \frac{1}{{\rm dim}}\,\,
2509: \lower0.25in\hbox{\epsfig{figure=cups.eps, height=0.5in}}
2510: \ee
2511: and
2512: \be\label{A-3}
2513: \int dg \,\,\, \lower0.25in\hbox{\epsfig{figure=3g.eps, height=0.5in}} =
2514: \lower0.25in\hbox{\epsfig{figure=clebsch.eps, height=0.5in}}
2515: \ee
2516: The intertwiner that is used in the last formula is chosen to be normalized so that:
2517: \be\label{A-theta}
2518: \lower0.35in\hbox{\epsfig{figure=theta.eps, height=0.7in}} = 1
2519: \ee
2520: whenever the 3 spins satisfy the triangular inequalities, and zero otherwise.
2521:
2522:
2523: \begin{thebibliography}{99}
2524:
2525: \bibitem{Barrett}
2526: J.~W.~Barrett,
2527: ``Feynman diagrams coupled to three-dimensional quantum gravity,''
2528: arXiv:gr-qc/0502048.
2529: %%CITATION = GR-QC 0502048;%%
2530:
2531:
2532: \bibitem{Freidel}
2533: L.~Freidel and E.~R.~Livine,
2534: ``Ponzano-Regge model revisited. III: Feynman diagrams and effective field
2535: theory,''
2536: arXiv:hep-th/0502106.
2537: %%CITATION = HEP-TH 0502106;%%
2538:
2539:
2540: \bibitem{PR} G. Ponzano and T. Regge, ``Semiclassical limit of Racah coefficients''
2541: in Spectroscopic and group theoretical methods in physics (Bloch ed.),
2542: North-Holland, 1968.
2543:
2544: \bibitem{Freidel-div}
2545: L.~Freidel and D.~Louapre,
2546: ``Diffeomorphisms and spin foam models,''
2547: Nucl.\ Phys.\ B {\bf 662}, 279 (2003)
2548: [arXiv:gr-qc/0212001].
2549: %%CITATION = GR-QC 0212001;%%
2550:
2551: %\cite{Freidel:2005ec}
2552: \bibitem{Majid-F}
2553: L.~Freidel and S.~Majid,
2554: ``Noncommutative harmonic analysis, sampling theory and the Duflo map in 2+1
2555: quantum gravity,''
2556: arXiv:hep-th/0601004.
2557:
2558: \bibitem{Freidel-CS}
2559: L.~Freidel and D.~Louapre,
2560: ``Ponzano-Regge model revisited. II: Equivalence with Chern-Simons,''
2561: arXiv:gr-qc/0410141.
2562:
2563: \bibitem{Barrett-Obs}
2564: J.~W.~Barrett, J.~M.~Garcia-Islas and J.~F.~Martins,
2565: ``Observables in the Turaev-Viro and Crane-Yetter models,''
2566: arXiv:math.qa/0411281.
2567: %%CITATION = MATH-QA 0411281;%%
2568:
2569: \bibitem{Boulatov}
2570: D.~V.~Boulatov,
2571: ``A Model of three-dimensional lattice gravity,''
2572: Mod.\ Phys.\ Lett.\ A {\bf 7}, 1629 (1992)
2573: [arXiv:hep-th/9202074].
2574: %%CITATION = HEP-TH 9202074;%%
2575:
2576:
2577: \bibitem{MM}
2578: P.~H.~Ginsparg and G.~W.~Moore,
2579: ``Lectures on 2-D gravity and 2-D string theory,''
2580: arXiv:hep-th/9304011.
2581: %%CITATION = HEP-TH 9304011;%%
2582:
2583:
2584:
2585: \bibitem{Sum}
2586: L.~Freidel and D.~Louapre,
2587: ``Non-perturbative summation over 3D discrete topologies,''
2588: Phys.\ Rev.\ D {\bf 68}, 104004 (2003)
2589: [arXiv:hep-th/0211026].
2590: %%CITATION = HEP-TH 0211026;%%
2591:
2592: \bibitem{BM}
2593: F.~A.~Bais and N.~M.~Muller,
2594: ``Topological field theory and the quantum double of SU(2),''
2595: Nucl.\ Phys.\ B {\bf 530}, 349 (1998)
2596: [arXiv:hep-th/9804130].
2597:
2598:
2599:
2600:
2601: %\cite{Mikovic:2001xi}
2602: \bibitem{Mikovic}
2603: A.~Mikovic,
2604: ``Spin foam models of matter coupled to gravity,''
2605: Class.\ Quant.\ Grav.\ {\bf 19}, 2335 (2002)
2606: [arXiv:hep-th/0108099].
2607: %%CITATION = HEP-TH 0108099;%%
2608:
2609:
2610: \bibitem{Laurent-recent}
2611: L.~Freidel,
2612: ``Group Field Theory: An overview,''
2613: arXiv:hep-th/0505016.
2614: %%CITATION = HEP-TH 0505016;%%
2615:
2616: \bibitem{Rovelli}
2617: M.~P.~Reisenberger and C.~Rovelli,
2618: ``Spacetime as a Feynman diagram: The connection formulation,''
2619: Class.\ Quant.\ Grav.\ {\bf 18}, 121 (2001)
2620: [arXiv:gr-qc/0002095].
2621: %%CITATION = GR-QC 0002095;%%
2622:
2623:
2624: \bibitem{Philippe} R.\ Dijkgraaf, V.\ Pasquier and Ph.\ Roche,
2625: ``Quasi-Hopf Algebras, Group Cohomology and Orbifold models,''
2626: Nucl.\ Phys.\ Proc Suppl.\ {\bf 18B},
2627: 60-72 (1990)
2628:
2629:
2630:
2631:
2632: \bibitem{Koorn}
2633: T.~H.~Koornwinder, B.~J.~Schroers, J.~K.~Slingerland and F.~A.~Bais,
2634: ``Fourier transform and the Verlinde formula for the quantum double of a
2635: finite group,''
2636: J.\ Phys.\ A {\bf 32}, 8539 (1999)
2637: [arXiv:math.qa/9904029].
2638: %%CITATION = MATH-QA 9904029;%%
2639:
2640:
2641: \bibitem{Roberts} J.~Roberts, ``Skein theory and Turaev-Viro invariants'',
2642: Topology {\bf 34} 771-787 (1995).
2643:
2644: \bibitem{Roberto}
2645: R.~De Pietri and C.~Petronio,
2646: ``Feynman diagrams of generalized matrix models and the associated manifolds
2647: in dimension 4,''
2648: J.\ Math.\ Phys.\ {\bf 41}, 6671 (2000)
2649: [arXiv:gr-qc/0004045].
2650: %%CITATION = GR-QC 0004045;%%
2651:
2652: \bibitem{Woronowicz} S. L. Woronowicz and S. Zakrzewski, ``Quantum
2653: deformations of the Lorentz group. The *-algebra level'', Compositio
2654: Math.\ {\bf 90} 211-243 (1994).
2655:
2656: \bibitem{Holography} K. Krasnov, ``Holography for the Lorentz group
2657: Racah coefficients'', Class.\ Quant.\ Grav.\ {\bf 22} 1933-1944 (2005).
2658:
2659: \end{thebibliography}
2660:
2661: \end{document}