0810.0771/ch3.tex
1: \chapter{Baby universes} \label{ch3}
2: 
3: Despite recent progress \cite{Ambjorn:2005db,Ambjorn:2005qt},
4: little is known about the ultimate
5: configuration space of quantum gravity on which its nonperturbative
6: dynamics takes place. This makes it difficult to decide
7: which (auxiliary)  configuration space to choose as starting
8: point for a quantization. In the context of a path integral
9: quantization of gravity, the relevant question is which class of
10: geometries one should be integrating over in the first place.
11: Setting aside the formidable difficulties in ``doing the integral", there is
12: a subtle balance between including too many geometries -- such
13: that the integral will simply fail to exist (nonperturbatively)  in any
14: meaningful way, even after renormalization -- and including too few
15: geometries, with the danger of not capturing a physically relevant part
16: of the configuration space.
17: 
18: A time-honored part of this discussion is the question of whether a
19: sum over different topologies should be included in the
20: gravitational path integral. The absence to date of a viable theory of
21: quantum gravity in four dimensions has not hindered speculation on
22: the potential physical significance of processes involving topology
23: change (for reviews, see \cite{Horowitz:1990qb,Dowker:2002hm}). Because such processes
24: necessarily violate causality, they are usually considered
25: in a Euclidean setting where the issue does not arise.
26: In our models for topology change we do wish to capture some Lorentzian aspects though.
27: In section \rf{subsec:Lorentzian aspects} we therefore address some of the causality issues
28: and argue that the causal structure of the manifolds is only modified mildly.
29: 
30: The focus of this chapter is devoted to the introduction of spatial topology change in
31: two dimensional quantum gravity. This might be viewed as a bit of an ironical undertaking since \emph{the} characteristic
32: that makes the quantum geometry of CDT better behaved than its Euclidean rival is precisely its fixed spatial topology!
33: We do however show that when a coupling constant is introduced the effect of the spatial topology changes is much less
34: severe. A natural scaling for this coupling constant is presented, where the dynamics of an individual spacetime
35: and splitting into baby universes both contribute to the continuum limit of the theory. Consequently, time scales canonically
36: and the limit where the coupling constant tends to zero is smooth and continuous and gives back the results
37: of the bare CDT model. This should be
38: contrasted with Euclidean quantum gravity defined through dynamical triangulations, since in that case
39: the continuum dynamics of the theory is completely dominated by the baby universes, making the geometry highly fractal.
40: 
41: Rather than going into the details of the theory with the coupling constant straightaway we will
42: first explain the role that baby universes play in the relation between Euclidean and Causal
43: dynamical triangulations in section \rf{sec:Euclidean results with causal methods}.
44: The formalism of CDT can easily be extended to allow for spatial topology
45: change whereby we admit the geometry to split into baby universes. It turns out that if one allows
46: for the baby universes the splitting process will dominate the path integral completely, wiping out
47: all traces of the dynamics of each individual spatial universe \cite{Ambjorn:1998xu}. At the same
48: time it is shown that this continuum limit corresponds uniquely to the dynamics of Euclidean
49: quantum gravity. Explicitly it is shown how to rederive the Hartle Hawking wavefunction and
50: propagator of Euclidean dynamical triangulations in the continuum limit. In this discussion we
51: closely follow \cite{Ambjorn:1999nc}.
52: 
53: In sections \rf{sec:Introducing the coupling constant},\rf{sec:Dynamics to all orders in the coupling}
54: and \rf{sec:Relation to random trees} we describe our published work where we generalize the above
55: described construction \cite{Ambjorn:2007xx}. Particularly, we associate a coupling constant that is reminiscent of
56: the string coupling constant with the spatial topology fluctuations.
57: We show that in a suitable scaling limit the spatial topology changes contribute to the path integral in a
58: controlled manner without dominating the quantum geometry.
59: 
60: \section{Euclidean results with causal methods} \label{sec:Euclidean results with causal methods}
61: 
62: In section \rf{subsec:Spatial topology change} we allow for spatial topology change
63: and show how to explicitly introduce the baby universes
64: in the discrete formalism and we compute the Hartle Hawking wavefunction in the continuum limit.
65: The result is the well known disc amplitude of two dimensional Euclidean quantum gravity.
66: 
67: \subsection{Spatial topology change} \label{subsec:Spatial topology change}
68: 
69: We now address the implementation of spatial topology changes by generalizing the discrete CDT framework.
70: Although a multitude of constructions that realize topology change exist, they are all equivalent in the continuum limit as
71: has been checked for some cases by the authors of \cite{Ambjorn:1998xu}.
72: In the following we introduce one specific realization and show that in the continuum limit the familiar
73: results from Euclidean quantum gravity are recovered.
74: 
75: 
76: The first step in the construction is to generalize the one step propagator, or transfer matrix,
77: of CDT by alleviating the constraint on the spatial topology of its initial loop.
78: Particularly, we include strips for which
79: the entrance loop, say of length $l_1$, has the topology of a ``figure eight''.
80: A natural procedure to create such a figure eight is to non-locally identify two points of a spatial universe with
81: topology of an $S^1$. Incorporating this pinching process leads to a factor of $l_1$ in the combinatorics for the one step
82: propagator since the pinching is allowed to happen at any of the $l_1$ vertices.
83: A baby universe is now created by associating one of the loops of the figure eight with the boundary of
84: a disc function whilst the other loop is associated with the initial loop of a regular one step propagator
85: (fig.~\ref{fig:dressedonesteppropagator}).
86: %Given that the disc function can be connected to both of the loops of the figure eight a factor of two
87: %needs to be introduced in the counting.
88: Combining all of the above, we can write the new transfer matrix in terms
89: of the old, or ``bare'',  transfer matrix and the as yet undetermined disc function
90: %
91: \beq \label{eq:onsteppropagatordressed}
92: G_\l(l_1,l_2;1)  = G_\l^{(b) }(l_1,l_2;1) + \sum_{l=1}^{l_1-1} l_1 w(l_1\mi l,g) \,G_\l^{(b) }(l,l_2;1),
93: \eeq
94: %
95: where it should be noted that we have dropped the superscript notation for the marking and have
96: used the notation $G_\l(l_1,l_2;1)=G^{(1,0)}_\l(l_1,l_2;1)$.
97: 
98: \begin{figure}[t]
99: \begin{center}
100: \includegraphics[width=4.3in]{dressedonesteppropagator}
101: \caption{Illustration of a one step propagator with a ``baby universe''.}
102: \label{fig:dressedonesteppropagator}
103: \end{center}
104: \end{figure}
105: 
106: This new, or ``dressed'', one step propagator satisfies the same composition law as the bare transfer matrix
107: \rf{eq:discrcompositionllmarked}, i.e.
108: %
109: \beq
110: G_\l(l_1,l_2;t_1+t_2)  = \sum_l G_\l(l_1,l;t_1) G_\l(l,l_2;t_2).
111: \eeq
112: %
113: Particularly this implies that the dressed one step propagator can still be used as a transfer matrix
114: despite the fact that the disc function contains an infinite sum over time steps
115: %
116: \beq\label{eq:discrcompositionlldressed}
117: G_\l(l_1,l_2;t)  = \sum_l G_\l(l_1,l;1) \;G_\l(l,l_2;t\mi 1).
118: \eeq
119: %
120: Performing a (discrete)  Laplace transformation of eq.~\rf{eq:discrcompositionlldressed}
121: leads to
122: \bea
123: {G(x,y;g;t)  =}~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~
124: ~~~~~~~~~~~~~~~~~~~~~~
125:  & &  \nonumber \\
126: \ointz  \left[G_\l^{(b) }(x,z^{-1};1) \pl x \frac{\prt}{\prt x}
127: \Bigl( w(x;g)  G_\l^{(b) }(x,z^{-1};1) \Bigr)  \right]  G(z,y;g;t \mi 1),
128: &&\nonumber \\
129: \!\!\!\!\!\!\!\! \label{eq:GxygtGdressed}
130: \eea
131: or, using the explicit form of the transfer matrix $G_{(b)}(x,z;g;1) $,
132: formula \rf{eq:G10xyg1},
133: \beq
134: G(x,y;g;t)  = \Bigl[1+x\frac{\prt w(x,g) }{\prt x}+ x w(x,g) \frac{\prt}{\prt x} \Bigr]
135: \, \frac{gx}{1-gx} \, G\Bigl( \frac{g}{1\mi gx},y;g;t\mi 1\Bigr) \label{eq:Gxygtdressed}.
136: \eeq
137: Note that this is not a closed equation, since so far neither the disc amplitude $w(x,g) $ nor $G(x,y;g;t) $ are known.
138: Although this means that we cannot derive the discrete expressions, it will be shown by using scaling arguments that
139: one can uniquely determine the continuum disc amplitude $W_\La(X) $ and propagator $G_\La(X,Y;T) $. As in the case
140: for CDT without topology change we assume canonical scaling for both the boundaries and the cosmological constant,
141: %
142: \beq
143: g=\frac{1}{2}e^{-a^2 \La},\quad x=e^{-a X},\quad y=e^{-a Y}.
144: \eeq
145: %
146: In the following arguments no specific choice for the scaling of the time variable will be assumed as its
147: scaling will be determined at a later stage. Interestingly, the composition law and the canonical scaling of the boundary lengths
148: are enough to determine the scaling of the dressed propagator
149: 
150: \beq
151: G_\l(l_1,l_2,t)  \buildrel{a\rightarrow0}\over\longrightarrow  a\, G_\La(L_1,L_2;T).
152: \eeq
153: 
154: \begin{figure}[t]
155: \begin{center}
156: \includegraphics[width=4in]{discfunctiondecompositionpure}
157: \caption{Decomposition of the marked CDT disc function in another CDT disc function and a propagator.}
158: \label{fig:discfunctiondecompositionpure}
159: \end{center}
160: \end{figure}
161: 
162: Changing the boundary conditions from fixed boundary length to fixed boundary cosmological constant
163: amounts to taking a Laplace transformation and implementing the canonical scaling of the boundary cosmological
164: constant $x=e^{-a X}$, leading to
165: 
166: \beq\label{top8}
167: G_\l(x,l_2,t)  \buildrel{a\rightarrow 0}\over\longrightarrow G_\La(X,L_2,T)
168: \eeq
169: 
170: and
171: 
172: \beq\label{top8x}
173: G_\l(x,y;t)  \buildrel{a\rightarrow 0}\over\longrightarrow a^{-1} G_\La(X,Y;T).
174: \eeq
175: 
176: The scaling relation for the disc function is not as simple to obtain as the scaling for the propagator
177: and we show that it actually depends on the scaling of time. Where for the propagator we use the composition law
178: to derive its scaling relation we use the following exact combinatorial identity to determine the scaling relation
179: for the disc function
180: 
181: 
182: \beq\label{an1}
183: g\ \frac{\prt w(x,g) }{\prt g} = \sum_t \sum_l G(x,l;g;t)  \, l\, w(l,g),
184: \eeq
185: or, after the usual Laplace transformation,
186: \beq\label{an2}
187: g\ \frac{\prt w(x,g) }{\prt g}=\sum_t \ointz G(x,z^{-1};g;t) \; \frac{\prt w(z,g) }{\prt z}.
188: \eeq
189: These identities reflect the fact that if one introduces a mark anywhere in the bulk by taking a derivative
190: with respect to $g$, the disc function can be decomposed into a propagator and another disc amplitude. The situation
191: for the bare model is illustrated in fig.~\ref{fig:discfunctiondecompositionpure}.
192: In the figure on the right we have highlighted the points that have a distance $t$ to the
193: entrance loop. Since the bare model does not allow for spatial topology change these points constitute one spatial universe
194: with the topology of an $S^1$. In fig.~\ref{fig:discfunctiondecompositiontop} the situation is illustrated in the case where one does allow for topology change,
195: in this case all points with equal distance $t$ from the entrance loop form a spatial section with the
196: topology of several $S^1$'s and the mark is located on one of them.
197: 
198: \begin{figure}[t]
199: \begin{center}
200: \includegraphics[width=4in]{discfunctiondecompositiontop}
201: \caption{Decomposition of the disc function including spatial topology change.}
202: \label{fig:discfunctiondecompositiontop}
203: \end{center}
204: \end{figure}
205: 
206: Since our current objective is to find the scaling for the disc function we assume the following general scaling ansatz
207: \beq\label{an3}
208: w(x,g) = w_{ns}(x,g) + a^{\eta}W_\La(X)  + \mbox{less singular terms}.
209: \eeq
210: In the case $\eta < 0$ the first term is irrelevant in the continuum limit, it does not appear in the computation
211: of any continuum quantities. However, if $\eta >0$  a term $w_{ns}$ will generically be present \cite{Ishibashi:1993sv}.
212: Additionally, the general ansatz for the scaling of the time variable reads as follows
213: %
214: \beq
215: T = a^{\ep}t,~~~~~\ep >0.
216: \eeq
217: %
218: As shown in section \rf{subsec:The continuum limit} both space and time scale canonically in the bare model which corresponds to $\ep=1$.
219: Below we show that allowing the branching into baby universes to contribute to the continuum limit
220: forces the scaling of time to be anomalous, creating an inherent asymmetry between the time- and space directions.
221: 
222: Inserting the scaling relations \rf{an3} and \rf{eq:scalingrelations} into eq.\ \rf{an2} we obtain
223: \bea
224: &&\frac{\prt w_{ns}}{\prt g}- 2a^{\eta-2}\frac{\prt W_\La(X) }{\prt \La} = \nonumber \\
225: &&\hspace{.6cm} \frac{1}{a^{\ep}}\int \d T \int dZ\;  G_\La(X,-Z;T) \bigg[ \frac{\prt w_{ns}}{\prt z} -a^{\eta-1}\frac{1}{z_{c}} \frac{\prt W_\La(Z) }{\prt Z}\bigg],
226: \label{an4}
227: \eea
228: where $(x,g) =(x_c,g_c) $ in the non-singular part.
229: 
230: From eq.\ \rf{an4} and the requirement $\epsilon >0$ it follows that the
231: only consistent choices for $\eta$ are
232: \begin{itemize}
233: \item[{\bf Scaling 1:}] {\boldmath{${\eta <\ }$}{\bf 0}}
234: 
235: As can be seen from \rf{an3}, this range of values corresponds to the situation where the non scaling part of the disc function
236: is irrelevant and the physics is completely independent of the cutoff,\
237: \beq\label{an5c}
238: a^{\eta-2} \frac{\prt W_\La(X) }{\prt \La} = \frac{a^{\eta-1}}{2 a^{\ep}}
239:   \int \d T \int dZ\;  G_\La(X,-Z;T)  \;\frac{1}{z_{c}}
240:   \frac{\prt W_\La(Z) }{\prt Z}.
241: \eeq
242: The continuum limit can be taken for any $\eta<0$ since \rf{an5c} does not depend on its explicit value. The value of
243: $\ep$ on the other hand is fixed and one needs to have $\ep=1$ for the continuum limit to exist.
244: Summarizing, if the scaling
245: of the disc function is such that non scaling contributions are negligible in the continuum limit, the time
246: variable automatically scales canonically. Obviously the bare CDT model falls in this class of scalings since it has
247: $\eta =-1$ and $\ep=1$.
248: 
249: \item[{\bf Scaling 2:}] {\bf 1}{\boldmath${\ < \eta < \ }$}{\bf 2}.
250: 
251: For this class of scalings for the disc function formula \rf{an4} splits into two equations
252: \beq\label{an5}
253: -a^{\eta-2} \frac{\prt W_\La(X) }{\prt \La} = \frac{1}{2 a^{\ep}}\,\frac{\prt w_{ns}}{\prt z}\bigg|_{z=x_c}\;\int \d T \int dZ\;  G_\La(X,-Z;T),
254: \eeq
255: and
256: \beq\label{an6}
257: \frac{\prt w_{ns}}{\prt g}\bigg|_{g=g_c} = -\frac{a^{\eta-1}}{a^{\ep}} \int \d T \int dZ\;  G_\La(X,-Z;T) \;\frac{1}{z_{c}} \frac{\prt W_\La(Z) }{\prt Z}.
258: \eeq
259: From \rf{an5} it follows that $a^{\eta-2}= \frac{1}{a^{\ep}}$ and from  \rf{an6} one sees that
260: $ \frac{a^{\eta-1}}{a^{\ep}}=1$. Combining these requirements we are led to the conclusion that $\ep= 1/2$ and $\eta=3/2$,
261: which are precisely the values found in Euclidean $2d$ gravity.
262: Let us further remark that eq.\ \rf{an5} in this case becomes
263: \beq\label{an5a}
264: -\frac{\prt W_\La(X) }{\prt \La} = \mbox{const.}\;G_\La(X,L_2=0).
265: \eeq
266: If one rescales the couplings it is possible in general to absorb the constant originating from the non universal terms.
267: This implies that the continuum dynamics of the theory strongly depends on the fact that there are non universal terms
268: surviving the continuum limit but their precise value does not play a pertinent role. It should be noted that \rf{an5}
269: expresses a different relation between the disc function and the propagator than was used for the bare model,
270:  it differs from \rf{eq:defwavefunctionLT} by a derivative with respect to the cosmological constant.
271: Finally, inserting $\ep= 1/2$ and $\eta=3/2$ into eq.\ \rf{an6} yields
272: \beq\label{an7}
273: \int \d T \int dZ\;  G_\La(X,-Z;T) \;\frac{\prt W_\La(Z) }{\prt Z} = \mbox{const},
274: \eeq
275: where as in \rf{an5a} the constant originates from the non scaling terms of the disc function and its value does not
276: play a significant role in the continuum theory.
277: \end{itemize}
278: 
279: 
280: \begin{figure}[t]
281: \begin{center}
282: \includegraphics[width=2in]{mini}
283: \caption{At every point in the quantum geometry there is an infinitesimal baby universe.}
284: \label{fig:mini}
285: \end{center}
286: \end{figure}
287: 
288: 
289: The relation \rf{an5a} possesses a remarkable interpretation in terms of baby universes. Basically it states that near
290: any mark in the bulk of the continuum Hartle Hawking wave function there is a baby universe at the scale of the cutoff.
291: Since the location of the mark is arbitrary it implies that
292: \emph{near every point there is a baby universe at the cutoff scale} which is illustrated in fig.~\ref{fig:mini}.
293: This does not mean however that all baby universes are of negligible size. An additional
294: indication that cutoff size geometries play an important role in the dynamics comes from the Laplace transform of \rf{an7}
295: %
296: \beq
297: \int \d T\int dL\;  G_\La(L,L';T) \;L'\;W_\La(L')  = \mbox{const}. \times ~\delta(L),
298: \eeq
299: %
300: which shows that the distribution of geometries is such that it is strongly peaked around universes that have minimal
301: boundary length. In the following we show that these microscopic artifacts
302: are an important feature of $2d$ Euclidean quantum gravity since the CDT model with baby universes exactly reproduces
303: the continuum equations of the Euclidean model in the scaling limit. Furthermore, to make contact with the
304: Euclidean theory, \rf{an5a} is used in an essential way.
305: 
306: Having derived the scaling relations we can now analyze the scaling limit of \rf{eq:Gxygtdressed} and find an equation
307: for the dynamics of the propagator. In order for the equation to have a scaling limit at all, $x_c,\, g_c$ and
308: $w_{ns}(x_c,g_c) $ must satisfy two relations which can be determined
309: straightforwardly from \rf{eq:Gxygtdressed}. The remaining continuum equation reads
310: \bea
311: a^\ep\frac{\prt}{\prt T}\, G_\La(X,Y;T) & =&-a \, \frac{\prt}{\prt X} \Bigl[ (X^2-\La)  G_\La(X,Y;T) \Bigr] \nn
312: &&-a^{\eta-1}\frac{\prt}{\prt X} \Bigl[W_\La(X)  G_\La(X,Y;T) \Bigr].
313: \label{top13}
314: \eea
315: The first term on the right-hand side of eq.\ \rf{top13} is precisely the one we have already encountered
316: in the bare model, while the second term is due to the creation of baby universes.
317: In case where $\eta \leq 1$ the first term in \rf{top13} is subdominant
318: and the scaling will not be compatible with baby universes in the continuum limit. So from \rf{top13} it can be seen that
319: for $\eta \leq 1$ we need to have that $\ep=1$ leaving us with the continuum differential equation for the
320: propagator of the bare model
321: %
322: \beq \label{eq:difequationGXYTpure}
323: \frac{\prt}{\prt T}\, G_\La(X,Y;T)  = - \frac{\prt}{\prt X} \Bigl[ (X^2-\La)  G_\La(X,Y;T) \Bigr],
324: \eeq
325: %
326: where we remind the reader that this differential equation can naturally be interpreted as a
327: Wick rotated Schr\"{o}dinger equation of a single string propagating with respect to its own time on world sheet,
328: %
329: \beq
330: \frac{\prt}{\prt T}\, G_\La(X,Y;T)  = \hat{H}_{X} G_\La(X,Y;T).
331: \eeq
332: %
333: From the Laplace transform of this equation one sees that $\hat{H}$ is a simple Hamiltonian containing a kinetic term,
334: a potential induced by the cosmological constant and no interaction terms,
335: %
336: \beq
337: \hat{H}^{\scriptscriptstyle marked}_L = -L\frac{\partial^2}{\partial L^2}+ \La \: L.
338: \eeq
339: %
340: For the second scaling however, where $ 1 < \eta < 2 $, the last term on the right-hand side of \rf{top13} will always
341: dominate over the first term. Consequently, the first term does \emph{not} survive the continuum limit leaving one with
342: an equation without a Hamiltonian containing a kinetic term. So for this scaling the dynamics of the world sheet is
343: governed purely by the interactions of splitting strings. Equivalently, we can say that
344: {\em once we allow for the creation of baby universes, this process will completely dominate the continuum limit.}
345: As we have seen from \rf{an5} and \rf{an6}, $(\eta,\ep) =(3/2,1/2) $ is the only consistent scaling that allows
346: for baby universes. Inserting
347: this scaling in \rf{top13} we obtain the following continuum equation
348: %
349: \beq\label{top16}
350: \frac{\prt}{\prt T}\, G_\La(X,Y;T)  = -\frac{\prt}{\prt X} \Bigl[W_\La(X)  G_\La(X,Y;T) \Bigr],
351: \eeq
352: %
353: %In addition we get $\ep = \eta-1$,
354: %in agreement with \rf{an6}. It follows that $\eta > 1$ and
355: %we conclude that $\ep=1/2,\eta=3/2$
356: %are the only possible scaling exponents if we allow for the creation
357: %of baby universes.
358: %These are precisely the scaling exponents obtained from
359: %two-dimensional Euclidean gravity in
360: %terms of dynamical triangulations, as we have already remarked. The topology
361: %changes of space have induced an anomalous dimension for $T$.
362: %If the second term on the right-hand side of \rf{top13} had been absent,
363: %this would have led to $\ep =1$, and the time $T$ scaling in the
364: %same way as the spatial length $L$.
365: %
366: %In summary, in the case $(\eta,\ep) =(3/2,1/2) $ eq.\ \rf{top13} leads
367: %to the continuum equation
368: %\beq\label{top16}
369: %\frac{\prt}{\prt T}\, G_\La(X,Y;T)  =
370: %-\frac{\prt}{\prt X} \Bigl[W_\La(X)  G_\La(X,Y;T) \Bigr],
371: %\eeq
372: which, combined with eq.\ \rf{an5a}, determines the continuum disc
373: amplitude $W_\La(X) $.
374: Integrating \rf{top16} with respect to $T$ and using that
375: $G_\La(L_1,L_2;T\equ 0) =\delta(L_1\mi L_2) $, i.e.\
376: \beq\label{top18a}
377: G_\La(X,L_2\equ 0;T\equ 0) =1,
378: \eeq
379: we obtain
380: \beq\label{top18}
381: -1 = \frac{\prt}{\prt X}\bigg[ W_\La(X)  \frac{\prt}{\prt\La} W_\La(X)\bigg].
382: \eeq
383: From dimensional analysis one can easily see that $W_\La^2(X)  = X^3 F(\SL/X)$. This implies that
384: the solution of the disc function reads as follows
385: \beq\label{top19}
386: W_\La(X)  = \sqrt{-2\La X + b^2 X^3+ c^2 \La^{3/2}}.
387: \eeq
388: However, not all values for $b$ and $c$ are physically acceptable.
389: The inverse Laplace transform of \rf{top19}, $W_\La(L)$, should be bounded for
390: all $L>0$. This constraint gives the following expression for the disc function
391: \beq\label{top20}
392: W_\La(X)  = b \Big(X-\frac{\sqrt{2}}{b\,\sqrt{3}} \,\SL\Big)
393: \sqrt{X+ \frac{2\sqrt{2}}{b\,\sqrt{3}}\SL},
394: \eeq
395: where the constant $b$ is a constant that reflects specific details of the discrete statistical model,
396: as described by equation \rf{an5a}. We discussed above that this constant does not have an obvious physical significance
397: since it can be absorbed into the cosmological constant.
398: Absorbing the irrelevant constant $b$ one obtains the disc function $W^{(eu) }_\La (X) $ of 2d Euclidean quantum gravity,
399: \beq\label{top21}
400: W_\La (X)  = (X -\oh \SL )  \sqrt{X+\SL}.
401: \eeq
402: Note that the defining equation for the disc function \rf{top16}, can alternatively be derived
403: by several methods within 2d Euclidean quantum gravity \cite{Kawai:1993cj,Ishibashi:1993sv,Watabiki:1993ym}.
404: In those computations it is clear that $T$ can be interpreted as the {\em geodesic distance} between
405: the initial and  final loop.
406: 
407: %Before showing that the anomalous scaling of the proper time --
408: %once baby universes are allowed --
409: %leads to an intrinsic fractal space-time dimension
410: %of four (rather than two), let us comment
411: %on the difference between the equations for
412: %the amplitudes \rf{an5c}-\rf{an6} for
413: %$(\eta,\ep)  =(-1,1) $ and $(\eta,\ep) =(3/2,1/2) $ respectively.
414: %In the first case there are no baby universes and
415: %eq.\ \rf{an5c} entails that only {\em macroscopic loops} at a
416: %distance $T$ from the entrance loop are important (as illustrated
417: %in Fig.\ \rf{identity}). On the other hand, the term $\prt W_\La(Z) /\prt Z$
418: %which describes the presence of these macroscopic loops is absent in
419: %eq.\ \rf{an5}.
420: %This is consistent with eq. \rf{an5a}, which shows explicitly
421: %that the length of the upper loop in Fig.\ \rf{identity} remains at
422: %the cut-off scale, and therefore can never become macroscopic.
423: %It is also consistent with the abundance of baby universes, since
424: %at any point in space-time the probability
425: %for creating a little ``tip'' of cut-off size
426: %will dominate. At the same time, the right-hand side of eq.\ \rf{an5c},
427: %that is, eq.\ \rf{an7}, will play no role when $1 < \eta <2$,
428: %being simply equal to a constant. This latter property is
429: %satisfied automatically, as can be seen by using an equation
430: %analogous to \rf{top16} for the exit instead of the entrance loop.
431: %Thus eq.\ \rf{an7} becomes proportional to
432: %\beq\label{last}
433: %\int_0^\infty \d T \frac{\prt}{\prt T} \;G(X,L_2 \equ 0;T)  =
434: %\mbox{const.},
435: %\eeq
436: %proving our previous assertion.
437: 
438: 
439: For a more detailed account on baby universes in
440: two dimensional Euclidean quantum gravity the reader is referred to \cite{Jain:1992bs}.
441: 
442: %
443: %\beq
444: %\frac{\partial}{\partial T} G^{\scriptscriptstyle{(1;0) }}_\La(X,Y;T) +\frac{\partial}{\partial X}\left[(X^2-\La) G^{(1;0) }_\La(X,Y;T) \right]=0
445: %\eeq
446: %
447: \section{Introducing the new coupling constant}\label{sec:Introducing the coupling constant}
448: 
449: In this section we present a model which can be viewed as a hybrid between Euclidean and causal quantum gravity. Recall
450: that the pure CDT model possesses a Hamiltonian that governs the dynamics of an individual string, but that there is no interaction term
451: in the dynamical equation to allow for spatial topology change \rf{eq:difequationGXYTpure}.
452: In the case of the Euclidean theory the opposite is true, the
453: equation governing the dynamics of the propagator only contains an interaction term and no kinetic or potential terms
454: for an individual string \rf{top16}. In the previous section it was shown that both theories can be obtained
455: by two different scaling
456: limits of one unifying statistical model based on CDT \rf{eq:onsteppropagatordressed}.
457: The current objective is to show that there is a natural adaptation of
458: this unifying model such that its scaling limit acquires single string dynamics while at the same time
459: allowing interactions in the form of spatial topology change. The essential ingredient in accomplishing this is to
460: introduce a new coupling in the statistical model \rf{eq:onsteppropagatordressed}
461: whose scaling is such that both the interaction term and the dynamical term
462: in \rf{eq:onsteppropagatordressed} contribute in the continuum limit.
463: %
464: \begin{figure}[t]
465: \begin{center}
466: \includegraphics[width=3in]{trousermodular}
467: \caption{A representation of the global aspects of a trouser geometry with modular parameters ($\tau_1, \tau_2, \tau_3$).}
468: \label{fig:trousermodular}
469: \end{center}
470: \end{figure}
471: %
472: Before discussing the construction of the model we examine some aspects of geometries with spatial topology change.
473: The simplest example of a two dimensional geometry with spatial topology change is the so-called ``trouser'' geometry.
474: A trouser geometry is a manifold with three boundary components with the topology of an $S^1$. In general the so-called
475: legs of the trousers can have any length which can be parameterized by three modular parameters as illustrated in
476: fig.~\ref{fig:trousermodular} .
477: 
478: We are interested however in a more restricted subclass of trouser geometries which are of the form depicted in
479: fig.~\ref{fig:trousercdt}.
480: We demand that we have an initial boundary, where every point on this boundary has the same distance to the two final loops.
481: This restriction means that on top of the length of the boundary components we need just two parameters to describe
482: the global characteristics of the geometries, one describing the length of the trouser leg belonging to the initial
483: boundary component and one that specifies the length of the trouser legs related to the final boundary components.
484: since we required the final boundary components to have the same distance to the initial boundary.
485: 
486: Now we ask ourselves the question, what is the contribution of such a geometry to the path integral? As described in
487: chapter \ref{ch2} the action for two dimensional gravity contains just a volume term proportional to the cosmological constant
488: and a topological term coming from the two dimensional Einstein-Hilbert action and the Gibbons-Hawking-York boundary term,
489: %
490: \beq
491:  S_M= \La V_M + \tfrac{2 \pi}{G_N}\chi(M),
492: \eeq
493: %
494: where $\chi(M)=\left(2-2\genus-b \right)$ is the Euler characteristic of the manifold.
495: The weight of an individual geometry in the path integral is therefore proportional to the exponential of the volume and the
496: exponential of the number of handles and boundary components.
497: %
498: \beq
499: G_{G_N,\Lambda}(L_1,L_2,T)  = \sum_{topol.} g^{2\genus}_S \int_M D[g_{\mu\nu}] e^{i \La V_M }.
500: \eeq
501: %
502: %
503: %\beq
504: %nietaf write PI ,\quad g_S = \exp{-1/G_N}
505: %\eeq
506: %
507: 
508: %
509: \begin{figure}[t]
510: \begin{center}
511: \includegraphics[width=4in]{trousercdt}
512: \caption{A representation of the global geometry of a trouser geometry in causal dynamical triangulations }
513: \label{fig:trousercdt}
514: \end{center}
515: \end{figure}
516: %
517: 
518: We see that geometries with the topology of a cylinder are not weighted by $g_S$, since the Euler characteristic of
519: a cylinder is zero. The Euler characteristic of a trouser geometry is one however, since the
520: trouser geometry has one more boundary component than the cylinder. Consequently the weight of the trouser geometry
521: in the path integral is not only determined by the cosmological constant, but also by the coupling $g_S$. For the purposes
522: of this thesis we are not particularly interested in computing amplitudes with the topology of a trouser, but
523: we are interested in calculating propagators, i.e.~amplitudes that have two macroscopic boundary components. If one
524: shrinks one of the final boundary components of a trouser geometry to zero one obtains a geometry with two
525: boundary components that we interpret as a ``cylinder geometry with one baby universe''.
526: Shrinking a boundary component does not alter the Euler characteristic, allowing us to conclude that a
527: geometry with $n$ baby universes contributes to the path integral with a $g_S^n$ topological weight. Note that in the
528: derivation of the disc amplitude of Euclidean dynamical triangulations of section
529: \rf{sec:Euclidean results with causal methods} we did not associate
530: a coupling constant to the baby universes. Subsequently, the formation of baby universe was not suppressed with the
531: result that the continuum limit is dominated by cutoff scale baby universes. In section
532: \rf{sec:Dynamics to all orders in the coupling} we show that adding the
533: coupling constant makes the term \emph{baby} universes a bit misleading since the outgrowths have a certain finite size
534: in the continuum limit and can be made arbitrarily big or small depending on the value of $g_S$.
535: 
536: \subsection{Lorentzian aspects} \label{subsec:Lorentzian aspects}
537: 
538: The reasoning in the previous section was based on a picture where the metrics are Wick rotated from Lorentzian to
539: Euclidean signature. Performing an inverse Wick rotation is not so easy however for geometries with baby universes,
540: since these geometries do not
541: admit Lorentzian metric everywhere. The reason for this
542: comes from the fact that it is not possible to find
543: a non vanishing vector field everywhere. It is however possible
544: to find a vector field everywhere except for a finite number of points.
545: For some values of the time parameter the universe develops a baby universe and the topology of a spatial universe
546: splits up into two different components. Precisely at
547: the moment of a split the spatial geometry has the topology of a figure eight and we see that the central point of the
548: figure eight is a saddle point if one views the geometry as being embedded in $\mathbb{R}^3$.
549: This point is called a Morse point in the
550: mathematically oriented literature \cite{Dowker:2002hm}.
551: Contrary to a generic point on the manifold, such a Morse point does not possess a unique
552: timelike vector field perpendicular to its spatial slice. In fact it has two future directed normal timelike vectors,
553: showing that the neighborhood of such a point has an anomalous causal structure. More precisely, such a point has
554: two future, and two past light cones which we refer to as a double light cone structure \cite{Dowker:2002hm}.
555: Consequently,
556: if one universe splits in two, each of the resulting universes carries a light cone belonging to the Morse point as
557: is illustrated in fig.~\ref{fig:morsepoint}.
558: Conversely, the universe carries the two past light cones of the Morse point before the split.
559: 
560: %
561: \begin{figure}[t]
562: \begin{center}
563: \includegraphics[width=4in]{morsepoint}
564: \caption{Illustration of the double light cone causal structure around a Morse point.}
565: \label{fig:morsepoint}
566: \end{center}
567: \end{figure}
568: %
569: 
570: Because of the peculiar causal structure around the Morse points we are not quite sure whether the usual definition of the
571: Wick rotation in CDT is also legitimate here. There are some indications that the Einstein Hilbert action develops complex
572: valued singularities at these points \cite{Louko:1995jw}. Although the Wick rotation at these points presents one with
573: an interesting conundrum, we confine ourselves to the Euclidean sector for the remainder of this thesis.
574: 
575: \section{Dynamics to all orders in the coupling}\label{sec:Dynamics to all orders in the coupling}
576: 
577: In this section we discuss the explicit construction of the model with a coupling constant for the spatial topology
578: changes and show that it can be solved to all orders in the coupling constant in a suitable continuum limit.
579: 
580: Our starting point is equation \rf{eq:onsteppropagatordressed} with the addition of a coupling to the interaction term,
581: where in section \rf{sec:Introducing the coupling constant} we argued
582: that the coupling can naturally be denoted as $g_s$,
583: %
584: \beq
585: G_\l(l_1,l_2;1)  = G_{\l}^{(b) }(l_1,l_2;1) + 2g_s \sum_{l=1}^{l_1-1} l_1 w_{\l,g_s}(l_1\mi l) \,G_\l^{(b) }(l,l_2;1).
586: \eeq
587: %
588: After a discrete Laplace transform one obtains
589: \bea
590: {G(x,y;g;t)  =}~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~
591: ~~~~~~~~~~~~~~~~~~~~~~
592:  & &  \nonumber \\
593: \ointz  \left[G^{(b) }(x,z^{-1};g;1) \pl ~2g_s~x \frac{\prt}{\prt x}
594: \Bigl( w_{g_s}(x;g)  G^{(b) }(x,z^{-1};g;1) \Bigr)  \right]  G(z,y;g;t \mi 1).
595: &&\nonumber  \label{eq:Gxygtdressedcomposition}\\
596: \!\!\!\!\!\!\!\!
597: \eea
598: If the baby universes are to survive the continuum limit the coupling constant in a controlled manner we have to scale $g_S$
599: with the lattice cutoff $a$. To find the appropriate scaling for the coupling we implement the following ansatz
600: %
601: \beq \label{eq:scalinggstr}
602: g_s = a^{\xi}g_S + \mbox{less singular terms.}
603: \eeq
604: %
605: Using the explicit form of the transfer matrix $G_\l^{(b) }(x,z^{-1};1) $ and inserting the scaling relations
606: \rf{eq:scalingrelations}, \rf{eq:scalinggstr} into
607: expression \rf{eq:Gxygtdressedcomposition} one is led to the following continuum equation,
608: %
609: \bea
610: a^\ep\frac{\prt}{\prt T}\, G_\La(X,Y;T) & =& -a \, \frac{\prt}{\prt X} \Bigl[ (X^2-\La)  G_\La(X,Y;T) \Bigr] \nn
611: &&- 2g_S \: a^{\eta+\xi-1}\frac{\prt}{\prt X} \Bigl[\cW_{\La,g_S}(X)  G_\La(X,Y;T) \Bigr].
612: \label{eq:semicontinuumdifequation}
613: \eea
614: %
615: If we demand that the first term on the right hand side of eq. \rf{eq:semicontinuumdifequation}
616: survives the continuum limit
617: we need $\eps =1$ as in the bare CDT model.
618: Contrary to \rf{top13} we have the additional freedom to scale the coupling constant $g_S$.
619: This enables us to adjust the scaling so that
620: also the second term survives the continuum limit yielding $\eta+\xi=2$. Inserting this relation in
621: \rf{eq:semicontinuumdifequation} we obtain
622: the desired dynamical equation that contains both the dynamical term for propagation of single strings and a term governing
623: string interactions,
624: %
625: \beq \label{eq:difequationGXYTdressed}
626: \frac{\partial}{\partial T}G^{(1;0) }(X,Y,T) =-\frac{\partial}{\partial X}\left[\left((X^2-\La) + 2 g_S \cW_{\La,g_S}(X) \right) G^{(1;0) }(X,Y,T) \right].
627: \eeq
628: %
629: We observe that this equation is of the same form as the equations for the CDT models without a coupling constant
630: \rf{eq:differentialequationGXYT},
631: %
632: \beq
633: \frac{\partial}{\partial T}G^{(1;0) }(X,Y;T) =-\frac{\partial}{\partial X}\left[\hcW_{\La,g_S}(X)  G^{(1;0) }(X,Y;T) \right],
634: \eeq
635: %
636: but with a different form of $\hcW(X) $,
637: %
638: \beq \label{eq:hcWintermsofW}
639: \hcW_{\La,g_S}(X) = (X^2-\La) \;+\;2 g_S\:\cW_{\La,g_S}(X).
640: \eeq
641: %
642: Note that requiring both terms in \rf{eq:difequationGXYTdressed} to survive the scaling limit does not fix
643: the scaling uniquely.
644: However, since we merely introduce a coupling constant to the baby universe model and did not introduce any new
645: configurations in the path integral, the disc function should still satisfy \rf{an1}.
646: As before we can use this relation to
647: further constrain the scaling. The dynamics of our new model requires $\ep=1$, so we are led to the conclusion that the
648: model falls into the first of the two scaling classes defined in section \rf{sec:Euclidean results with causal methods}
649: implying $\eta<0$. Consequently, the relation
650: between the continuum disc function and the continuum propagator is the same as for the bare CDT model,
651: %
652: \beq \label{eq:contmarkinglawW}
653: \frac{\partial\cW_{\La,g_S}(X) }{\partial \La} = \int dT \int d Z G^{(1;0) }(X,-Z;T) \frac{\partial\cW_{\La,g_S}}{\partial Z}.
654: \eeq
655: %
656: It seems that we are still left with a predicament since we are unable to specify the exact scaling for the disc function.
657: In fact it is only an illusory problem, since the disc function only appears in the dynamical equations in combination
658: with the coupling constant as $g_S \cW$. Therefore, we in principle only need to determine the scaling of this
659: combination which means that the relevant part of the scaling is already determined by the dynamical equation itself.
660: In the model we consider however, the scaling is completely fixed by requiring the
661: disc function to reduce to the disc function of the bare CDT model for zero $g_S$, implying
662: %
663: %
664: \beq
665: w_{\l,g_s}(x) = a^{-1}\cW_{\La,g_S}(X)  , \quad g_s = a^{3}g_S.
666: \eeq
667: %
668: %
669: In the remainder of this section \rf{eq:contmarkinglawW} and \rf{eq:difequationGXYTdressed}
670: are used to derive a differential equation for $\hcW_{\La,g_S}(X) $.
671: The equation will be rather implicit however, since the equation does not only depend on $\hcW_{\La,g_S}(X)$, but it also
672: depends explicitly on the solution of the equation $\hcW_{\La,g_S}(X) =0$. Remarkably, one is
673: able to solve the equation uniquely provided the disc functions $\cW_{\La,g_S}(L) $ satisfy the natural physical
674: requirement that they fall of at infinity. As a first step we solve the dynamical equation
675: \rf{eq:difequationGXYTdressed} to obtain
676: the propagator in terms of $\hcW_{\La,g_S}(X)$,
677: %
678: \beq
679: G^{(1;0) }(X,Y,T) =\frac{\hcW_{\La,g_S}(\bX(T) ) }{\hcW_{\La,g_S}(X) }\frac{1}{\bX(T) +Y}.
680: \eeq
681: %
682: Inserting this into the consistency condition \rf{eq:contmarkinglawW} and performing the integration over $Z$ we obtain
683: %
684: \beq \label{eq:ddlaWisintdT}
685: \frac{\partial\cW_{\La,g_S}(X) }{\partial \La} = -\int d T \frac{\hcW_{\La,g_S}(\bX(T) ) }{\hcW_{\La,g_S}(X) }\frac{\partial\cW_{\La,g_S}}{\partial \bX(T) }.
686: \eeq
687: %
688: To evaluate the integral we conveniently use the characteristic equation and convert the integral over time into an integral
689: over $\bX$,
690: %
691: \beq \label{eq:characteristicequationhcW}
692: \frac{d\bX}{dT}=-\hcW_{\La,g_S}(\bX(T)).
693: \eeq
694: %
695: %The simplest example to illustrate this change of variables is the definition of the time variable itself in terms of
696: %$\hcW_{\La,g_S}(X) $.
697: %By separating variables one can integrate the characteristic equation and express the relation between the
698: %time variable and $\hcW_{\La,g_S}(X) $,
699: %
700: %\beq
701: %T\:=\:\int \limits^{\infty}_{0} d T = \int\limits^{\bX_{\infty} }_{X} d\bX \frac{1}{\hcW_{\La,g_S}(\bX)},
702: %\eeq
703: %
704: %where $\bX_{\infty} =\bX(T=\infty) $ is the solution of $\hcW_{\La,g_S}(\bX_{\infty}) =0$.
705: Applying the characteristic equation to \rf{eq:ddlaWisintdT} we obtain
706: %
707: \beq
708: \frac{\partial\cW_{\La,g_S}(X) }{\partial \La} =\frac{1}{\hcW_{\La,g_S}(X) }
709: \int\limits^{\bX_{\infty} }_X d\bX \frac{\partial\cW_{\La,g_S}}{\partial \bX},
710: \eeq
711: where one can easily evaluate the integral since the integrand is a total derivative,
712: %
713: \beq \frac{\partial\cW_{\La,g_S}(X) }{\partial \La}=\frac{\cW_{\La,g_S}(X)-\cW_{\La,g_S}(\bX_{\infty}) }{\hcW_{\La,g_S}(X) }. \eeq
714: %
715: This equation can be rewritten as an equation for $\hcW_{\La,g_S}(X) $ by reexpressing the disc functions as
716: %
717: \beq
718: \cW_{\La,g_S}(X) = -\frac{\hcW_{\La,g_S}(X) -(X^2-\La)  }{2 g_S},
719: \eeq
720: %
721: giving
722: %
723: \beq
724: \frac{\partial\hcW_{\La,g_S}(X) }{\partial\La}=\frac{\bX_{\infty}^2-X^2}{\hcW_{\La,g_S}(X) },
725: \eeq
726: %
727: or equivalently,
728: %
729: \beq \label{eq:difequationhcW}
730: \frac{\partial\hcW_{\La,g_S}(X) ^2}{\partial \La}= 2(\bX_{\infty}^2-X^2).
731: \eeq
732: %
733: This is the defining equation for $\hcW_{\La,g_S}(X) $ that we intended to derive. The simple form of the equation is deceptive
734: as $\bX_{\infty}$ is in fact a completely unknown function of $g_S$ and $\La$, the only thing we know is that it is defined as
735: the solution of $\hcW_{\La,g_S}(X) =0$. Another unknown arises when we try to solve \rf{eq:difequationhcW}
736: by integrating both sides with
737: respect to $\La$, the integration ``constant'', $C(g_S,X) $, can in principle be a general function of $g_S$ and $X$. So we
738: conclude that \rf{eq:difequationhcW} only determines $\hcW_{\La,g_S}(X) $ up to two functions.
739: Explicitly, the integral of \rf{eq:difequationhcW} can be written as follows,
740: %
741: \beq \label{eq:hcWsqintermsoffandg}
742: \hcW_{\La,g_S}(X)^2=\La^2 f\left(\textstyle{\frac{g_S}{\sqrt{\La}^3}}\right) -2 X^2\La + X^4h\left(\textstyle{\frac{g_S}{X^3}}\right),
743: \eeq
744: %
745: where we have used dimensional analysis to parameterize the integration constant and the integral of $\bX_{\infty}$ by two
746: dimensionless functions,
747: %
748: \beq
749: \La^2 f\left(\textstyle{\frac{g_S}{\sqrt{\La}^3}}\right)  = 2\int^{\La} d\La' \bX_{\infty}\left(\textstyle{g_S,\sqrt{\La'}^3}\right),
750: \quad C(g_S,X) = X^4h\left(\textstyle{\frac{g_S}{X^3}}\right).
751: \eeq
752: %
753: Below we use the following series expansions
754: %
755: \beq
756: f \left(\textstyle{\frac{g_S}{\sqrt{\La}^3}}\right) =
757: \sum\limits_{n=0}^{\infty}f_n\left(\textstyle{\frac{g_S}{\sqrt{\La}^3}}\right) ^n,
758: \quad h \left(\textstyle{\frac{g_S}{X^3}}\right) =
759: \sum\limits_{n=0}^{\infty}h_n\left(\textstyle{\frac{g_S}{X^3}}\right) ^n.
760: \eeq
761: %
762: %into \rf{eq:hcWsqintermsoffandg} to obtain
763: %
764: %\beq \label{eq:defhcWalltaylorcoef}
765: %\hcW_{\La,g_S}(X) ^2=(X^2-\La) ^2+\sum\limits_{n=1}^{\infty}g_S^n\left(h_n X^{4-3n}+f_n \sqrt{\La}^{4-3n}\right).
766: %\eeq
767: %
768: We show that the lowest order coefficients of the expansions are determined by requiring consistency
769: with the bare CDT model. Amazingly, \emph{all} higher order coefficients are
770: uniquely determined by demanding that the inverse Laplace transform of the disc function $\cW(X) $ falls of at infinity
771: at arbitrary order in the $g_S$ expansion. From the definition of $\hcW(X) $ in the pure model \rf{eq:defhcWpure}
772: one readily sees that to lowest order in $g_S$ we need to have
773: %
774: \beq
775: \hcW_{\La,g_S}(X)^2=X^4 -2X^2\La +\La^2 + \cO(g_S),
776: \eeq
777: %
778: which implies $h_0=1,f_0 =1$ when comparing to \rf{eq:hcWsqintermsoffandg}.
779: Using this result we write the disc function as follows,
780: %
781: \beq \label{cWtaylor}
782: \cW_{\La,g_S}(X) =\frac{-(X^2-\La) +(X^2-\La) \sqrt{1+\frac{\sum\limits_{n=1}^{\infty}g_S^n\left(h_n X^{4-3n}+f_n \sqrt{\La}^{4-3n}\right) }{(X^2-\La) ^2}}}{2g_S}.
783: \eeq
784: %
785: If we expand this expression to lowest order in $g_S$ we obtain
786: %
787: \beq \label{cWlowestorder}
788: \cW_{\La,g_S}(X) =\frac{1}{4}\frac{h_1 X +f_1\sqrt{\La}}{X^2-\La}+\cO(g_S).
789: \eeq
790: %
791: By equating this result to the marked disc function of the bare model
792: %
793: \beq \label{cWpure}
794: \cW_{\La,0}(X) = \frac{1}{X+\sqrt{\La}},
795: \eeq
796: %
797: we conclude that $h_1=4,f_1=-4$.
798: To obtain the higher order coefficients we proceed by analyzing the $g_S$ expansion of the disc function order by order.
799: The first order correction to the disc function is obtained by inserting $h_1=4,f_1=-4$ into \rf{cWtaylor} and expanding
800: to first order in the coupling,
801: %
802: \beq
803: \cW_{\La,g_S}(X) =\textstyle{\frac{1}{X+\sqrt{\La}}+g_S\left(-\frac{1}{(X^2-\La) (X+\sqrt{\La}) ^2}+\frac{h_2
804: (4X) ^{-2}+f_2(4\sqrt{\La}) ^{-2}}{X^2-\La}\right) +\cO(g_S^2)}.
805: \eeq
806: %
807: After an inverse Laplace transform we obtain
808: %
809: \beq
810: \sqrt{\La}^3\cW_{\La,1}(L) =\frac{1}{8} e^{L \sqrt{\Lambda }}
811: (f_2+h_2-1) -\frac{h_2 L}{4} +\frac{1}{8} e^{-L\sqrt{\Lambda}}
812: \left(2 \Lambda  L^2+2 \sqrt{\Lambda } L-f_2-h_2+1\right),
813: \eeq
814: %
815: where we have introduced the following notation
816: %
817: \beq
818: \cW_{\La,g_S}(X) =\sum\limits_{n=0}^{\infty}\cW_{\La,n}(X) g_S^n.
819: \eeq
820: %
821: Demanding the disc function to fall of at infinity implies that the terms proportional to
822: $L$ and $e^{L \sqrt{\Lambda}}$ must vanish, leading to $h_2=0,f_2=1$. So the result
823: for the disc function at first order in the coupling reads as follows,
824: %
825: \beq \label{cWpowergsone}
826: \cW_{\La,1}(L) =\frac{e^{-L \sqrt{\Lambda }} L \left(\sqrt{\Lambda} L+1\right) }{4 \Lambda},
827: \eeq
828: %
829: which can be confirmed by explicitly computing the ``Feynman diagram'' where the spatial universe
830: is allowed to split once. Obtaining the higher order coefficients is a bit messy, but the iterative procedure
831: to compute them is completely analogous to the calculation for $h_2$ and $f_2$. At each order of the $g_S$ expansion
832: the disc function has the same form as \rf{cWpowergsone}. Every $\cW_{\La,n}(L) $ contains three terms,
833: a polynomial term in $L$, a term proportional to $e^{\sqrt{\La}L}$ and a term proportional to $e^{-\sqrt{\La}L}$.
834: Demanding the disc function to be bounded at infinity implies that the polynomial and $ \cO (e^{\sqrt{\La}L})$
835: terms should vanish.
836: If one additionally uses the known results for $h_{n-1}$ and $f_{n-1}$ one obtains the $h_n$ and $f_n$ coefficients
837: uniquely. It turns out that boundedness of the disc function is such a stringent constraint that we need
838: $h_{n}=0$ for all $n\geq2$. Inserting the nonzero coefficients of $h\left(\textstyle{\frac{g_S}{X^3}}\right) $
839: into \rf{eq:hcWsqintermsoffandg} we obtain
840: %
841: \beq
842: \hcW_{\La,g_S}(X) ^2=X^4 -2X^2\La + 4 g_S X + F(g_S,\La),
843: \eeq
844: %
845: where
846: %
847: \beq
848: F(g_S,\La) = \La^2 f(q),\quad q=\textstyle{\frac{g_S}{\sqrt{\La}^3}}.
849: \eeq
850: As noted above, demanding the disc function to be bounded at infinity also fixes all $f_n$.
851: Contrary to the $h_n$ coefficients however, the $f_n$ coefficients are rather nontrivial but can be computed by the
852: algorithm sketched above.
853: Employing a symbolic computer program such as Mathematica one can readily compute the first coefficients ($\sim 20$).
854: Making use of Sloan's database of integer sequences it is possible to find a closed analytical expression
855: for the coefficients $f_n$ in terms of Euler gamma functions,
856: %
857: \beq
858: f_0 = 1,\quad f_n=\frac{\Gamma(\frac{3}{2}n-2) }{\Gamma(\frac{1}{2}n+1) \Gamma(n) },\quad n\geq1.
859: \eeq
860: %
861: Given these coefficients we can sum the Taylor expansion and obtain the full non perturbative result for $f\left(q\right) $
862: %
863: \beq
864: f(q) = {\textstyle \frac{2}{3}+\frac{1}{3}}\, _2F_1\left({\textstyle-\frac{1}{3},-\frac{2}{3};\frac{1}{2}; \frac{2}{3 \sqrt{3}}}q\right) -
865: {\textstyle 4_2F_1}\left({\textstyle-\frac{1}{6},\frac{1}{6};\frac{3}{2}; \frac{2}{3 \sqrt{3}}}q \right).
866: \eeq
867: %
868: This means that we have solved our model and we can present the disc function with
869: a non perturbative sum over spatial topologies,
870: %
871: \beq \label{eq:cWallorders}
872: \cW_{\La,g_S}(X) =\frac{-(X^2-\La) + \sqrt{X^4 -2X^2\La + 4 q \sqrt{\La}^3  X + \La^2 f\left(q\right) }}{2g_S}.
873: \eeq
874: %
875: %Introducing dimensionless variables gives,
876: %
877: %\beq
878: %{\omega}(x) =\frac{-(x^2-1) + \sqrt{x^4 -2x^2 + 4 q x + f\left(q\right) }}{2q }
879: %\eeq
880: %
881: 
882: 
883: %
884: %\beq
885: %\frac{\partial \cW(L) }{\partial \La} = \int dT \int d L'
886: %G^{(1;0) }(L,L',T) L'\cW(L')
887: %\eeq
888: %
889: %\beq
890: %\frac{\partial\cW(X) }{\partial
891: %\La}=\frac{\cW(X) +\frac{1}{2g_S}(X_{\infty}^2-\La) }{\hcW(X) }
892: %\eeq
893: %
894: %\beq
895: %\frac{\partial}{\partial
896: %\La}(\hcW(X) -(X^2-\La) ) =\frac{\hcW(X) +(X_{\infty}^2-X^2) }{\hcW(X) }
897: %\eeq
898: %
899: %\beq \hcW(X) ^2=2\int d\La X_{\infty}\left(\scriptstyle{g_S,\sqrt{\La}^3}\right)  -2 X^2\La + C(g_S,X)
900: %\eeq
901: %
902: %\beq
903: %\hcW(X) ^2=(X^2-\La) ^2+X^4\sum\limits_{n=1}^{\infty}h_n\left(\textstyle{\frac{g_S}{X^3}}\right) ^n+
904: %\sum\limits_{n=1}^{\infty}f_n\left(\textstyle{\frac{g_S}{\sqrt{\La}^3}}\right) ^n
905: %\eeq
906: %
907: %
908: 
909: 
910: 
911: \section{Relation to random trees} \label{sec:Relation to random trees}
912: 
913: In this section we give an alternative derivation of the disc function dressed with spatial topology changes
914: \rf{eq:cWallorders}
915: to show robustness of the result and to give more insight into the details of the quantum geometry. Particularly,
916: the derivation we present below highlights the random tree structure of the configurations. To make the connection as clear
917: as possible we start by deriving the one point function for a random tree model. Sometimes this one point function for
918: random trees is referred to as a partition function for rooted branched polymers. The random tree model we
919: consider is a statistical
920: model consisting of edges that are weighted with a fugacity $z$ and three-valent vertices with a coupling constant
921: $\lambda$. A convenient way to view the partition function $w(z,g) $ is that it is the generating function for the number of
922: random trees,
923: %
924: \beq
925: w(z,\lambda)  = z \sum \limits_{m=0}^{\infty} w_{2m} \left(\lambda z^2\right) ^m,
926: \eeq
927: %
928: where we use the fact that every vertex except the initial, or marked, vertex is accompanied by two edges.
929: An easy way to evaluate the partition function $w(z,\lambda) $ is to notice that it should solve the following equation
930: %
931: \beq \label{branchedpolymereq}
932: w(z,\lambda)  = z + \lambda z w(z,\lambda) ^2.
933: \eeq
934: %
935: 
936: %
937: \begin{figure}[t]
938: \begin{center}
939: \includegraphics[width=3in]{branchedpolymer}
940: \caption{Pictorial representation of the iterative equation for the one-point function $w(\lambda,z) $ for branched polymers.}
941: \label{fig:branchedpolymer}
942: \end{center}
943: \end{figure}
944: %
945: 
946: The interpretation of this equation is illustrated in fig.~\ref{fig:branchedpolymer}.
947: Equation \rf{branchedpolymereq} has two solutions but only
948: one solution is compatible with the initial condition that there is only one tree with one edge, i.e.~$w_{2m} = 1$.
949: This solution is given by
950: %
951: \beq
952: w(z,\lambda)  = \frac{1-\sqrt{1-4 \lambda z^2}}{2 \lambda z},
953: \eeq
954: %
955: which has the following series expansion,
956: %
957: \beq
958: w(z,\lambda)  = z \sum \limits_{m=0}^{\infty} \frac{(2m) !}{m!(m+1) !} \left(\lambda z^2\right) ^m.
959: \eeq
960: %
961: So we see that the number of random trees with $m$ edges, $w_{2m}$, is given by the $m^{th}$ Catalan number.
962: This random tree model can be regarded as a simple model possessing some features of $\phi^3$ theory. Indeed
963: it is possible to formulate scalar $\phi^3$ theory in $\mathbb{R}^d$ as a sum over connected diagrams in the same
964: way as the random tree model. One merely needs to replace the fugacity of the edges $z$ by the standard scalar propagators.
965: Here we view the random trees not as a model for particles interacting in an ambient space, but as a simple model
966: for a theory of ``interacting universes'' in two dimensional quantum gravity. The idea is similar to the basic
967: idea behind string theory, we ``blow up'' the Feynman diagrams by replacing the propagators of an interacting field
968: theory by the propagators of a string theory as is illustrated in fig.~\ref{fig:stringinteraction}.
969: In our case this means that we replace
970: the rather trivial propagators of the branched polymer model $z$ by propagators that we computed from CDT \rf{eq:G10XYT}.
971: We remind the reader that the resulting model can either be viewed as a toy model for quantum gravity with
972: topology change or as a string theory without a target space.
973: To make the analogy as close as possible we write the equation for the generating function of the branched polymer as follows,
974: 
975: %
976: \begin{figure}[t]
977: \begin{center}
978: \includegraphics[width=3in]{stringinteraction}
979: \caption{Illustration of spatial topology change as a string like generalization of a particle interaction.}
980: \label{fig:stringinteraction}
981: \end{center}
982: \end{figure}
983: %
984: 
985: %
986: \beq \label{eq:branchedpolymereqwG}
987: w(z,\lambda)  = w_0(z)  + \lambda G_0(z)  w(z,\lambda) ^2,
988: \eeq
989: %
990: where $w_0(z) $ and $G_0(z) $ denote the one point function and the two point function respectively, to zeroth order
991: in the coupling $\lambda$. From \rf{branchedpolymereq} we observe that the expressions for these objects coincide
992: $w_0(z)=G_0(z)=z$. This is an understandable
993: coincidence, since the branched polymer is a very simple model. The
994: equation analogous to \rf{eq:branchedpolymereqwG} for our model of interacting spatial universes based on CDT is
995: %
996: \beq \label{eq:iterationequationbabyuniverses}
997: \cW_{\La,g_S}(L) = \cW_{\La,0}(L)  + g_S \int d T dL_1 dL_2 G^{(1,1) }_{\La,0}(L,L_1+L_2;T) \cW_{\La,g_S}(L_1) \cW_{\La,g_S}(L_2).
998: \eeq
999: %
1000: This is the defining equation for the disc function $\cW(L)$.
1001: It is similar to the way equation \rf{eq:branchedpolymereqwG} defines
1002: the one point function of the branched polymers $w(z,\lambda)$,
1003: since the disc function is the CDT analogue of the branched polymer one point function $w(z,\lambda) $.
1004: Instead of solving \rf{eq:iterationequationbabyuniverses} we work with its Laplace transformed analogue,
1005: %
1006: \beq \label{eq:iterationequationbabyuniversesXY}
1007: \cW_{\La,g_S}(X) = \cW_{\La,0}(X)  + g_S \int dZ_1 dZ_2 G^{(1,1}_{\La,0}(X;-Z_1,-Z_2) \cW_{\La,g_S}(Z_1) \cW_{\La,g_S}(Z_2),
1008: \eeq
1009: %
1010: where the propagator is given by
1011: %
1012: \beq
1013: G^{(1,1) }(X;Y_1,Y_2;T) =
1014: \frac{\hcW_{\La,0}(\bX) }{\hcW_{\La,0}(X) }\left[\frac{1}{(\bX+Y_1) ^2(\bX+Y_2) }+\frac{1}{(\bX+Y_1) (\bX+Y_2) ^2}\right].
1015: \eeq
1016: % `
1017: Performing the integrations over $Z_1$ and $Z_2$ in \rf{eq:iterationequationbabyuniversesXY} gives
1018: %
1019: \beq
1020: \cW_{\La,g_S}(X) = \cW_{\La,0}(X)  + g_S \int dT  \frac{\hcW_{\La,0}(\bX(T) ) }{\hcW_{\La,0}(X) }\frac{\partial \cW_{\La,g_S}(\bX) ^2}{\partial \bX}.
1021: \eeq
1022: %
1023: As in \rf{eq:ddlaWisintdT} it is convenient to use the characteristic equation to convert the
1024: integral over $T$ into an integral over $\bX(X,T) $,
1025: %
1026: \beq
1027: \cW_{\La,g_S}(X) = \cW_{\La,0}(X)  + \frac{g_S}{\hcW_{\La,0}(X) } \int\limits_X^{\bX_{\infty}}d\bX \frac{\partial \cW_{\La,g_S}(\bX) ^2}{\partial \bX}.
1028: \eeq
1029: %
1030: Seeing that the integrand is a total derivative one easily obtains
1031: %
1032: \beq
1033: \cW_{\La,g_S}(X) = \cW_{\La,0}(X)  + g_S \frac{\cW_{\La,g_S}(\bX_{\infty}) ^2-\cW_{\La,g_S}(X) ^2}{\hcW_{\La,0}(X)}.
1034: \eeq
1035: %
1036: Since this is a second order polynomial equation for $\cW(X) $ it is readily solved and we obtain,
1037: %
1038: \beq
1039:  \cW_{\La,g_S}(X) =\frac{-\hcW_{\La,0}(X) + \hcW_{\La,g_S}(X) }{2 g_S},
1040: \eeq
1041: %
1042: where
1043: %
1044: \beq \label{eq:hcWxbinfty}
1045:  \hcW_{\La,g_S}(X) =\sqrt{\hcW_{\La,0}(X) ^2+4 g_S\left(\hcW_{\La,0}(X) \cW_{\La,0}(X) + g_S \cW_{\La,g_S}(\bX_{\infty}) ^2\,\right)}.
1046: \eeq
1047: %
1048: Note that this equation was derived independently of the particular form of $\cW_{\La,0}(X) $ and $\hcW_{\La,0}(X) $. Currently, however
1049: we are interested in an interacting model based on CDT. So we require the disc function and the propagator to reduce
1050: to the results obtained in the bare CDT theory,
1051: %
1052: \beq
1053: \cW_{\La,0}(X)  = \frac{1}{X+\sqrt{\La}},\quad \hcW_{\La,0}(X)  = X^2-\La,\quad \bX_{\infty} = \sqrt{\La}.
1054: \eeq
1055: %
1056: Inserting into \rf{eq:hcWxbinfty} gives,
1057: %
1058: \beq
1059: \hcW_{\La,g_S}(X) ^2=(X^2-\La) ^2+4 g_S\left((X-\sqrt{\La}) + g_S \cW_{\La,g_S}(\sqrt{\La}) ^2\,\right),
1060: \eeq
1061: %
1062: which can be conveniently written as
1063: %
1064: \beq \label{eq:hcWsq}
1065: \hcW_{\La,g_S}(X) ^2=X^4 -2X^2\La + 4 g_S  X + F(g_S,\La),
1066: \eeq
1067: %
1068: where
1069: %
1070: \beq
1071: F(g_S,\La) = \La^2-4g_S\sqrt{\La}+4g_S^2\cW_{\La,g_S}(\sqrt{\La}) ^2.
1072: \eeq
1073: %
1074: Given \rf{eq:hcWsq} and \rf{eq:hcWintermsofW} we can obtain the following result for the disc function,
1075: %
1076: \beq
1077: \cW_{\La,g_S}(X) =\frac{-(X^2-\La) + \sqrt{X^4 -2X^2\La + 4 g_S  X + F(g_S,\La) }}{2g_S},
1078: \eeq
1079: %
1080: which is exactly of the same form as derived in the previous section \rf{sec:Dynamics to all orders in the coupling}
1081: but so far we have not yet
1082: determined the precise form of $F(g_S,\La) $. To abbreviate the notation,
1083: we can convert the equation to dimensionless units by
1084: dividing all dimensional quantities by appropriate powers of the cosmological constant
1085: %
1086: \beq
1087: x = \tfrac{X}{\sqrt{\La}},~~Y=\tfrac{Y}{\sqrt{\La}},~~ q = \tfrac{g_S}{\sqrt{\La}^3},
1088: \eeq
1089: %
1090: and obtain
1091: %
1092: \beq
1093: \omega_q(x) =\frac{-(x^2-1) +\hat{\omega}_q(x) }{2q},
1094: \eeq
1095: %
1096: with
1097: %
1098: \beq \label{eq:omegaxfq}
1099: \hat{\omega}_q(x) =\sqrt{x^4 -2x^2 + 4 q x  + f(q)}.
1100: \eeq
1101: %
1102: In the derivation of section \rf{sec:Dynamics to all orders in the coupling} the disc function is expanded
1103: in powers of $q$ and the Taylor coefficients of $F(q)$ are uniquely determined by demanding that the
1104: disc functions decay at infinity at every order of the expansion. Below we take a different route to obtain
1105: the $f(q) $ by again using the characteristic equation in an essential way. Surprisingly, we find $f(q)$
1106: in a form that appears very different from the results of section \rf{sec:Dynamics to all orders in the coupling},
1107: but is in fact exactly the same.
1108: Let us recall that the characteristic equation \rf{eq:characteristicequationhcW}
1109: can be used to define the time variable in terms of $\hat{\omega}_q(x)$ by
1110: %
1111: \beq
1112: t = \int\limits_{\bx_{\infty} }^{x}\frac{d\bx}{\hat{\omega}(\bx)}.
1113: \eeq
1114: %
1115: Notice that we can only integrate $t$ all the way to infinity if $\hat{\omega}_q(x)$ has a simple zero since $\bx_{\infty}$ is
1116: defined as the solution of $\hat{\omega}_q(x)=0$. Together with \rf{eq:omegaxfq} it implies that $\hat{\omega}_q(x)=0$
1117: should be of the following form,
1118: %
1119: \beq \label{eq:omegaxccc}
1120: \hat{\omega}_q(x) =(x-c) \sqrt{(x+c_{+}) (x+c_{-}) },
1121: \eeq
1122: %
1123: where $c,c_{+}$ and $c_{-}$ are all functions of $q$ that we determine below by taking the square of \rf{eq:omegaxccc}
1124: and equating
1125: it with the square of \rf{eq:omegaxfq}. This gives us four equations for the four unknown functions
1126: $c(q),c_{+}(q),c_{-}(q) $ and $f(q) $,
1127: one equation for each power of $x$, enabling one to solve the system completely.
1128: If one solves the equation belonging to $x^3$
1129: one can eliminate one function and we can write
1130: %
1131: \beq \label{eq:omegaxcu}
1132: \hat{\omega}_q(x) =(x-c) ^2(x+c+\sqrt{u}) (x+c-\sqrt{u}),
1133: \eeq
1134: %
1135: where $c_{+}= c + \sqrt{u}$ and  $c_{-}= c-\sqrt{u}$.
1136: If we now expand equation \rf{eq:omegaxcu} in powers of $x$ and than equate it to the square of \rf{eq:omegaxfq}
1137: we obtain
1138: %
1139: \beq
1140: x^4 - (2c^2+u)  x^2+ 2cu x + c^2(c^2-u) =x^4 - 2x^2 + 4 q x  + f.
1141: \eeq
1142: %
1143: From this we extract three equations, two simple relations expressing
1144: $f$ and $u$ in terms of $c$,
1145: %
1146: \beq \label{eq:fintermsofc}
1147: f = c^2(3c^2-2),
1148: \eeq
1149: %
1150: \beq \label{eq:uintermsofc}
1151: u = 2 - 2c^2,
1152: \eeq
1153: %
1154: and a third order polynomial equation for $c$
1155: %
1156: \beq \label{eq:thirorderequationcq}
1157: c^3-c = -q.
1158: \eeq
1159: %
1160: Inserting the solution of the polynomial equation for $c$ in \rf{eq:fintermsofc} gives $f(q) $ which together
1161: with \rf{eq:omegaxfq} allows us to find the complete solution for $\hat{\omega}(x) $.
1162: The solution of \rf{eq:thirorderequationcq} is most conveniently expressed in terms of $\tilde{q}= \frac{2}{3\sqrt{3}}\,q$
1163: as follows,
1164: %
1165: \beq \label{eq:cintermsofz}
1166: \sqrt{3}\,c(\tilde{q}) =z_q+\frac{1}{z_q},
1167: \eeq
1168: %
1169: where
1170: %
1171: \beq \label{eq:zintermsofq}
1172: z_q=\left(-\tilde{q}+\sqrt{\tilde{q}^2-1}\right) ^{\frac{1}{3}}.
1173: \eeq
1174: %
1175: Combining \rf{eq:cintermsofz}, \rf{eq:zintermsofq} and \rf{eq:fintermsofc} we obtain,
1176: %
1177: \beq
1178: f(q) =\frac{1}{3}\left( z_q^4 + 2 z_q^2 + 2 + \frac{2}{z_q^2} +\frac{1}{z_q^4}\right),
1179: \eeq
1180: %
1181: which appears to be very different from the expression found in the previous section,
1182: %
1183: \beq
1184: f(q) = {\textstyle \frac{2}{3}+\frac{1}{3}}\, _2F_1\left({\textstyle-\frac{1}{3},-\frac{2}{3};\frac{1}{2}; \frac{2}{3 \sqrt{3}}}q\right) -
1185: {\textstyle 4_2F_1}\left({\textstyle-\frac{1}{6},\frac{1}{6};\frac{3}{2}; \frac{2}{3 \sqrt{3}}}q \right),
1186: \eeq
1187: %
1188: but when one compares the Taylor expansion of both expressions we see that they are fully equivalent.
1189: 
1190: \section{Summary}
1191: 
1192: To set the stage for our model we reviewed the known relation between Euclidean and causal quantum gravity
1193: defined by dynamical triangulations in section \rf{sec:Euclidean results with causal methods}.
1194: Some results of Euclidean quantum gravity can be derived by generalizing
1195: the formalism of causal dynamical triangulations to allow for spatial topology change. No ``energy penalty'' is
1196: associated with these topological fluctuations, manifesting itself by the fact that infinitesimal
1197: baby universes dominate the path integral. The fractal nature of the quantum geometry is reflected by the
1198: non canonical values for both the ``time'' variable and the Hausdorff dimension.
1199: 
1200: In section \rf{sec:Introducing the coupling constant} we introduced a coupling constant for the spatial topology
1201: changes. From the Einstein Hilbert action of an elementary manifold with a change of spatial topology,
1202: the trouser geometry,
1203: we argued the naturalness of such a coupling. Additionally, we examined the geometry around the point where
1204: the spatial topology change occurs, the Morse point, and recalled that the causal structure around such a
1205: point is non standard. Particularly, the Morse point features a doubling of the light cone structure, it
1206: possesses two past and two future light cones.
1207: 
1208: In section \rf{sec:Dynamics to all orders in the coupling} we discussed the details of the construction
1209: our new model of two dimensional quantum gravity with spatial topology change.
1210: The discrete kinematical structure of the model is similar to the introductory
1211: section \rf{sec:Euclidean results with causal methods}, the continuum behavior on the other hand
1212: is completely different. The presence of the coupling allows us to define a continuum limit where
1213: manifolds with all spatial topologies contribute but complicated topologies are suppressed by powers
1214: of the coupling constant. Especially, we were able to derive the disc function to all orders in the coupling
1215: constant and sum the series uniquely!
1216: 
1217: An alternative derivation of the disc function dressed with topology fluctuations was presented in
1218: section \rf{sec:Relation to random trees}. We showed that the disc function of the model can be derived
1219: from an iterative equation that is very similar to the generating function equation that defines the
1220: one point function of rooted random trees, or equivalently branched polymers. Besides providing additional
1221: insight into the structure of the quantum geometry of the model we also found that the hypergeometric functions
1222: that appeared in the result of \rf{sec:Dynamics to all orders in the coupling} can be written in terms of a
1223: solution of a third order polynomial equation.
1224: 
1225: \begin{comment}
1226: %
1227: \beq
1228: W(L) = W_{CDT}(L)  + g_S \int dL_1 dL_2
1229: G(L,L_1+L_2) W(L_1) W(L_2)
1230: \eeq
1231: %
1232: \beq
1233: W^{(1) }(X) = W_{CDT}^{(1) }(X)  + g_S \int dL_1 dL_2G^{(1,1) }(X,L_1+L_2) W^{(1) }(L_1) W^{(1) }(L_2)
1234: \eeq
1235: %
1236: \beq
1237: G^{(1,1) }(X,L_1+L_2;T) =
1238: \frac{\hcW_0(\bX(T) ) }{\hcW_0(X) }(L_1+L_2) e^{-\bX(T) (L_1+L_2) }
1239: \eeq
1240: %
1241: %
1242: \beq
1243: G^{(1,1) }(X,Y_1,Y_2;T) =
1244: \frac{\hcW_0(\bX(T) ) }{\hcW_0(X) }\left[\frac{1}{(\bX(T) +Y_1) ^2(\bX(T) +Y_2) }+\frac{1}{(\bX(T) +Y_1) (\bX(T) +Y_2) ^2}\right]
1245: \eeq
1246: %
1247: \beq
1248: G^{(1,1) }(X,Y_1,Y_2;T) =
1249: \frac{\hcW_0(\bX_T) }{\hcW_0(X) }\left[\frac{1}{(\bX_T+Y_1) ^2(\bX_T+Y_2) }+\frac{1}{(\bX_T+Y_1) (\bX_T+Y_2) ^2}\right]
1250: \eeq
1251: %
1252: \beq
1253: \cW(X) = \cW_0(X)  + g_S \int dT \int dZ_1 dZ_2 \frac{\hcW_0(\bX(T) ) }{\hcW_0(X) }\frac{2}{(Z_1-\bX(T) ) ^2(Z_2-\bX(T) ) }\cW(Z_1) \cW(Z_2)
1254: \eeq
1255: %
1256: \beq
1257:   g_S\cW(X) ^2 + \hcW_0(X) \cW(X)  -\hcW_0(X) \cW_0(X)  - g_S \cW(\bX_{\infty}) ^2=0
1258: \eeq
1259: %
1260: \beq
1261: % \cW(X) =\frac{-\hcW_0(X) +\sqrt{\hcW_0(X) ^2+4 g_S\left(\hcW_0(X) \cW_0(X) + g_S \cW(\bX_{\infty}) \right) }}{2 g_S}
1262: \eeq
1263: 
1264: 
1265: \beq
1266: G^{(1,1) }(X,L_1+L_2) =
1267: \frac{e^{-\sqrt{\La}(L_1+L_2) }-e^{-X(L_1+L_2) }}{X^2-\La}
1268: \eeq
1269: %
1270: 
1271: \beq
1272: W^{(1) }(X) = \frac{1}{X+\sqrt{\La}} + g_S \int dZ_1 dZ_2
1273: G^{(1,1) }(X;-Z_1,-Z_2) W^{(1) }(Z_1) W^{(1) }(Z_2)
1274:  \eeq
1275: %
1276: \beq
1277: G^{(1,1) }(X;Z_1,Z_2) =\frac{1}{X^2-\La}\left(\frac{1}{(Z_1-\sqrt{\La}) (Z_2-\sqrt{\La}) }-\frac{1}{(Z_1-X) (Z_2-X) }\right)
1278: \eeq
1279: %
1280: \beq W^{(1) }(X) =\frac{1}{X+\sqrt{\La}}
1281: +\frac{g_S}{X^2-\La}\left(W^{(1) }(\sqrt{\La}) ^2-W^{(1) }(X) ^2) \right)
1282: \eeq
1283: %
1284: \beq \frac{q}{x^2-1}\cW(x) ^2+\cW(x) -\left(\frac{1}{x+1}+\frac{g_S
1285: \cW(1) ^2}{x^2-1}\right) =0 \eeq
1286: %
1287: \beq 2g_S \cW(x) = -(x^2-1) \pm\sqrt{(x^2-1) ^2+4q(x-1) +4q^2\cW(1) ^2}
1288: \eeq
1289: %
1290: \beq \hcW(X) =\sqrt{x^4-2x^2+4qx+b(q)  } \eeq
1291: %
1292: \beq b(q) =1-4q+4q^2\cW(1) ^2 \eeq
1293: %
1294: \beq \hcW(x) =(x-c(q) ) \sqrt{(x+c_{+}(q) ) (x-c_{-}(q) ) } \eeq
1295: %
1296: \beq \hcW(x) ^2 = x^4 - (2c^2+u) x^2+2cux+c^2(c^2-u)  \eeq
1297: %
1298: \beq u=2-2c^2,u=\frac{2q}{c},b(q) =c^2(c^2-u)  \eeq
1299: %
1300: \beq -c^3+c=q \eeq
1301: %
1302: \beq c_+(q) =c+\sqrt{u(q) }=c(q) +\sqrt{2-2 c(q) ^2} \eeq
1303: %
1304: \beq c^{-}_q=c_q-\sqrt{u_q}=c_q-\sqrt{2-2 c_q^2} \eeq
1305: %
1306: \beq b_q=c_q^2(c_q^2-u) =-c_q^2(c_q^2+2)  \eeq
1307: %
1308: 
1309: At this point is is interesting to make a small digression from the derivation of the disc function to examine
1310: the limit $X\rightarrow0, \bX(X,T)  \rightarrow0$. In this limit ... reduces to
1311: %
1312: \beq
1313: \cW(0) = \cW_0(0)  + \frac{g_S}{\hcW_0(0) }\cW(0) ^2
1314: \eeq
1315: %
1316: %Using
1317: %
1318: %\bea
1319: %\cW_0(0)  &=& lim \frac{1}{X+\sqrt{\La}}= \frac{1}{\sqrt{\La}}} \\
1320: %\hcW_0(0) &=& lim X^2-\La = -\La
1321: %\eea
1322: %
1323: If we now restore the dependence on $g_S$ and $\La$ in the notation by replacing $\cW(0)  \rightarrow \cW(g_S,\La) $
1324: and use the known expressions for $\cW_0(X) $ and $\hcW_0(X) $ we obtain
1325: %
1326: \beq
1327: \cW(g_S,\La) =\frac{1}{\sqrt{\La}} + \frac{g_S}{\La}\cW(g_S,\La) ^2
1328: \eeq
1329: %
1330: \end{comment}
1331: