hep-th0505104/hh.tex
1: %\documentclass[11pt]{article}
2: \documentclass{JHEP3}
3: \usepackage{amsmath,amsfonts,bbm,euscript, amssymb}
4: %\usepackage{epstopdf}
5: %\DeclareGraphicsRule{.tif}{png}{.png}{`convert #1 `dirname #1`/`basename #1 .tif`.png}
6: \usepackage{cite}
7: \usepackage{epsfig}
8: %\usepackage{geometry}
9: \usepackage{graphicx}
10: \usepackage{psfrag}
11: %\tolerance=10000
12: %\documentstyle[preprint,aps,epsfig, multicol]{revtex}
13: %\tighten
14: %\draft
15: %\widetext
16: %\usepackage {epsfig}
17: %\input{epsf.sty}
18: %\preprint{}
19: %\newcommand{\nc}{\newcommand}
20: \def\al{\alpha}
21: \def\om{\omega}
22: \def\R{\sideset{^3}R}
23: \def\be{\begin{equation}}
24: \def\ee{\end{equation}}
25: \def\baray{\begin{eqnarray}}
26: \def\earay{\end{eqnarray}}
27: \def\ba{\begin{eqnarray}}
28: \def\ea{\end{eqnarray}}
29: \def\V{{\bf V}}
30: \def\lsim{\roughly<}
31: \def\gsim{\roughly>}
32: \def\Dbar{\overline{D}}
33: \def\ap{{\alpha^{\prime}}}
34: \def\A{\hat A}
35: \def\Z{\hat Z}
36: \def\hf{\frac12}
37: \def\kp{k_\perp}
38: \def\kpl{k_\parallel}
39: \def\I{\boldmath{I}}
40: \def\F{\cal{F}}
41: \newcommand{\rb}{\bar{\rho}}
42: \newcommand{\p}{\phi}
43: \newcommand{\pb}{\bar{\phi}}
44: \newcommand{\cb}{\bar{c}}
45: \newcommand{\s}{\sigma}
46: \newcommand{\Sb}{\bar{S}}
47: \newcommand{\kb}{\bar{\kappa}}
48: \newcommand{\Sp}{S^{\prime}}
49: %\def\Y{\mathcal{Y}}
50: 
51: 
52: 
53: \title{The Boundedness of Euclidean Gravity and the Wavefunction of the Universe
54: %Hartle-Hawking Wavefunction and the Boundedness of Gravity
55: }
56: \author{Sash Sarangi and S.-H. Henry Tye \\
57:  Laboratory for Elementary Particle Physics, Cornell
58: University, Ithaca, NY 14853 \\
59: E-mail:\email{sash@lepp.cornell.edu},
60: \email{tye@lepp.cornell.edu}}
61: 
62: %\date{\today}
63: 
64: \abstract{When the semi-positive cosmological constant is dynamical, the naive Euclidean Einstein 
65: action is unbounded from below and the Hartle-Hawking wavefunction of the universe is not normalizable. With the inclusion of back-reaction (a crucial point), the presence of the metric 
66: perturbative modes (as well as matter fields) as a radiation term is introduced by quantum fluctuation. They act as the environment (that is, to be integrated or traced out), and 
67: introduce a correction term that provides a bound to the Euclidean action. As a result,  the improved wavefunction is normalizable. That is, decoherence plays an essential role in the  consistency of quantum gravity. In the spontaneous creation of the universe, this improved 
68: wavefunction allows one to compare the tunneling probabilities from absolute nothing (i.e., not even 
69: classical spacetime) to various vacua (with different large spatial dimensions and different low 
70: energy spectra) in the stringy cosmic landscape.
71: %instanton with higher multipoles that are perturbations around a pure deSitter instanton with a scalar field. We examine the consequences of such coupling on the tunnelling rate.
72: }
73: 
74: \begin{document}
75: 
76: \tableofcontents
77: 
78: \section{Introduction}
79: 
80: In the probing of the origin of our universe, a particularly attractive idea is the tunneling from 
81: absolute nothing (here, nothing means not even classical spacetime) \cite{Vilenkin:1982de}, or 
82: equivalently, the Hartle-Hawking (HH) no-boundary wave function of the universe \cite{Hartle:1983ai}.
83: The basic idea is illustrated in Figure 1. 
84: \begin{figure}
85: \begin{center}
86: \epsfig{file=s_4.eps, width=9cm}
87: \vspace{0.1in}
88: \caption{Tunneling from absolute nothing to a deSitter or Inflationary Universe.}
89: \label{fig1}
90: \end{center}
91: \end{figure}
92: It is strongly believed that, in superstring theory, there is a vast number of stable and 
93: meta-stable vacua, with up to 9 or 10 large spatial dimensions. This is the cosmic landscape.
94: If we can reliably calculate the tunneling probability from absolute nothing 
95: %(that is, no classical spacetime)
96: to any point in this vast landscape, one may argue that the origin of our universe should be 
97: the point in the landscape with the largest tunneling probability. 
98: This tunneling probability is given by (the absolute square of) the wavefunction of the universe. However, the HH wavefunction is not normalizable. 
99: Furthermore, it gives an answer that contradicts observations. It is well-known that the 
100: problem lies in the infrared/macroscopic limit, so quantum/stringy corrections will not be helpful. 
101: In Ref\cite{Firouzjahi:2004mx}, we conjectured that a normalizable wavefunction results if 
102: decoherence effect is included. One can then apply this improved wavefunction to find the preferred vacuum in the cosmic landscape (the one emerging from tunneling from nothing with the largest tunneling probability), which then evolves to today's universe. We call this proposal the Selection of the Original Universe Principle (SOUP); that is, today's universe must lie along the road that starts with the 
103: original preferred vacuum, arguably the 4-dimensional inflationary universe supported by observations. In this paper, we investigate more carefully the decoherence effect on the 
104: tunneling probability/wavefunction. As expected, the result supports the basic underpinning of SOUP, though the details are somewhat more involved. 
105:  
106: Consider the simple Einstein theory in 4 dimensions with a positive cosmological constant :
107: \ba
108: S_{E} = - \frac{1}{16 \pi G}\int d^4 x \sqrt{-g} \left( R - 2\Lambda\right)
109: \ea
110: The tunneling probability from nothing to a closed deSitter (or inflationary) universe with 
111: cosmological constant $\Lambda$ is given by
112: \ba
113: \label{EactionG}
114: \Gamma = |\Psi|^2\simeq e^{-S_E}  \quad \quad S_E=- \frac{3 \pi}{G\Lambda} 
115: \ea
116: where $S_E$ is the Euclidean action of the $S^4$ instanton \cite{Gibbons:1978ac}.
117: Suppose $\Lambda$ is dynamical, as in a model with 4-form field strengths \cite{Brown:1987dd}. 
118: The tunneling probability $\Gamma$ seems to allow us 
119: to pick out the universe with the largest probability. However, the Euclidean action $S_E$ is 
120: unbounded from below as $\Lambda \rightarrow 0$, so $\Gamma \to \infty$.
121: This has at least 2 obvious problems :
122: 
123: (1) The wavefunction $\Psi \simeq e^{3 \pi/2G\Lambda}$ is not normalizable. 
124: %Unitarity can be satisfied if $I_E$ is bounded from below. If the above formula is correct, 
125: %there is no normalization that can satisfy unitarity. 
126: Furthermore, one can easily show that there are other topological instantons with even 
127: more negative Euclidean action and so larger (actually, infinitely larger)  tunneling 
128: probability \cite{Coleman:1988tj}. This implies the presence of an inconsistency.
129: 
130: (2) Phenomenologically, since $\Gamma \rightarrow \infty$ as  $\Lambda \rightarrow 0$, 
131: it will imply the preference of tunneling to a flat universe with zero $\Lambda$, which 
132: contradicts the big bang history of our universe. 
133: Since the size of the universe, the cosmic scale factor $a, \sim 1/\sqrt{\Lambda}$, this will imply 
134: that the tunneling is to a universe of super-macroscopic size, contradicting one's intuition
135: that tunneling is a quantum process and so should be microscopic.
136:  
137: This issue is an outstanding problem since the early 1980s. 
138: The inconsistency has prevented the proper application of the whole idea.  
139: Possible resolution to this problem has been suggested (see 
140: \cite{Linde:1998gs} and references therein):
141: \begin{itemize}
142: \item
143: Instead of the usual $t \rightarrow - i \tau$, 
144: one may choose instead to rotate time to Euclidean time via 
145: $t \rightarrow i \tau$ \cite{Linde:1984mx}. This may work for pure gravity, but the 
146: inclusion of a scalar field (e.g., as required for inflation) 
147: leads to catastrophic consequences, since 
148: the scalar field theory becomes unbounded from below \cite{Rubakov:1984bh}. 
149: \item
150: One may argue that this problem will be corrected by quantum 
151: corrections or string theory corrections. However, it is easy 
152: to see that this is unlikely to be the case. Note that the 
153: problem occurs for small $\Lambda$, or large universe, since 
154: the cosmic scale factor $a \sim 1/\sqrt{\Lambda}$.
155: So this is more like an infrared or macroscopic problem than an ultra-violet problem. 
156: In fact, one can easily see that the loop correction \cite{Hawking:1976ja,Gibbons:1978ji}
157: does not solve the problem. Also, recent work \cite{Ooguri:2005vr}, where the exact HH wavefunction
158: is obtained in topological string theory, it seems that 
159: $\Lambda \to 0$ is again preferred, just like the original HH wavefunction.
160: 
161: \item This last property, namely, that the unboundedness problem 
162: becomes acute when the deSitter universe becomes large, or 
163: macroscopic (actually super-macroscopic), naturally suggests that 
164: the resolution should lie in decoherence \cite{Firouzjahi:2004mx}.
165: \end{itemize}
166: 
167: In Ref.\cite{Firouzjahi:2004mx}, we argue that the mini-superspace formulation is inadequate, 
168: and we propose that the inclusion of decoherence effects due to other modes provides a lower bound to $S_E$. We then speculate how the improved 
169: wavefunction may be used to select the stringy vacuum with the largest tunneling probability
170: from absolute nothing.
171: In this paper, we shall show that decoherence indeed provides a bound to the Euclidean gravity
172: action, though the formula for the tunneling probability may be more involved.  
173: Here, we shall focus on the pure 4-D Einstein gravity case. Generalization to more dimensions is straightforward and will be briefly discussed.
174: 
175: %add this
176: The basic idea of the approach is widely used in physics. Given any complicated problem, we usually
177: follow only a limited set of degrees of freedom, called the system. The remaining degrees 
178: of freedom, called the environment, are either ignored, or, in a better approximation, integrated out. 
179: In decoherence, integrating out the environment can cause the quantum system to behave like
180: a classical system. In effective field theory in particle physics, the massive modes are integrated 
181: out to produce higher 
182: dimensional operators (interaction terms) for the light modes. A famous example is the integrating
183: out of the W and Z bosons in the electroweak theory that yields the 4-Fermi weak interactions. Another example is the integrating out of the hidden sector in supergravity phenomenology. (Note that this has nothing to do with loop corrections.)
184: In the Wilson approach in quantum field theory, where high momentum modes are integrated out, and in the above cases, it is clear that the physics may crucially depend on the effects coming from integrating out the unobserved modes; that is, they cannot be ignored. 
185: %
186: %Integrating out the environment provides an effective bound to the resulting theory can be seen in a simple example. 
187: %
188: %For an example more closely related to the problem at hand, consider a somewhat artificial model of 2 scalar fields $\phi$ and $\eta$, with an effective potential $V= -m^2\phi^2/2 + M^2\eta^2/2 + \lambda \phi^2 \eta$. Here $\phi$ may have a tachyonic mass. Either $\eta$ has a large tachyonic mass (technically, we should treat it as an auxiliary field without a kinetic term) or $\lambda$ is imaginary.  At low energy, if we ignore $\eta$ completely, we find that the effective potential $V(\phi)$ is unbounded from below. However, integrating out the heavy mode $\eta$ introduces an effective interaction term $|\lambda|^2 \phi^4/M^2$, which provides a lower bound to the resulting effective potential $V(\phi)$. 
189: %
190: %The situation in Euclidean gravity is quite similar.
191: Of course, loop corrections can also induce dynamics/interactions that are not present in the tree level, as for example in the Coleman-Weinberg model, in light-light scattering, and in the running of couplings. In QCD, the running of the coupling emerges from the renormalization group improved quantum corrections, not just a normal 1-loop effect.
192: %however, this is not our main concern here.
193: We argue that this is also the situation in the study of the wavefunction of the universe. One may consider our result as a back-reaction improved quantum correction, not just a normal 1-loop effect.
194: 
195: The basic idea applied to tunneling is quite simple. It is well-known that the quantum tunneling of a particle with mass $M$, or the system, is suppressed if it interacts with an environment.
196: Consider a particle at $q=0$ in the potential $V(q)$ as shown in Figure 2.
197: \begin{figure}
198: \begin{center}
199: \epsfig{file=quartic.eps, width=8cm}
200: \vspace{0.1in}
201: \caption{Potential $V(q)$ with a metastable minimum at
202: $q=0$, and a stable minimum at $q = q_{+}$.}
203: \label{fig2}
204: \end{center}
205: \end{figure}
206: %The WKB approximation is good provided that the height of the barrier is larger than 
207: %$\om_0$, where $$M \om_0^2 =  \frac{\partial^2 V}{\partial q^2}|_{q=0}$$
208: In the WKB approximation, its tunneling rate is given by 
209: \ba
210: \Gamma &\simeq&  \exp (-S_0) \nonumber \\
211: S_0 &=& \int^{q_0}_0 \sqrt{2MV(q)} dq
212: \ea
213: where $S_0$ is the Euclidean action of the bounce, i.e., the instanton 
214: solution \cite{Coleman:1977py}.
215: Note that, for $V(q)$ bounded from below, $S_0$ is bounded 
216: from below, as required by consistency.
217: 
218: In a more realistic situation, the particle interacts with a set of other particles, or modes, 
219: say $x_{\al}$. However, we are only interested in the quantum status of $q$, so these other modes 
220: are integrated out in the path integral, or traced over in the density matrix formulation. 
221: They provide the environment. Their presence typically introduces a frictional force to 
222: the evolution of $q$.  It was shown by Sethna \cite{Sethna:1981dr} and by Caldeira and Leggett \cite{Caldeira:1982uj}, that the bounce $S_0$ increases to
223: $S \simeq S_0 \left( 1 + {\hat \eta}\right)$
224: where $\hat \eta > 0$ is proportional to the coefficient of friction (see Appendix A for details). 
225: That is, the interaction with the environment suppresses the tunneling rate. (This suppression takes place even if no friction is generated.) One may understand this result in a 
226: number of (equivalent) ways :
227: \begin{itemize}
228: \item 
229: As a quantum system, 
230: \ba
231: S &=& \int  \sqrt{2MV(q,x_{\al})} ds \nonumber \\
232: ds^2 &=& dq^2 + \sum \frac{m_{\al}}{M} d x_{\al}^2
233: \ea
234: That is, the increase in $S$ is due entirely to the longer path 
235: length in the many dimensional $(q, x_{\al})$ space.
236: \item 
237: The interaction of $q$ with the $x_{\al}$ interferes with its 
238: attempt in tunneling. One may view the interaction with $x_{\al}$ 
239: as attempts to observe $q$. Repeated measurements of $q$ or 
240: repeated attempts of measuring $q$ suppresses the tunneling rate.
241: This is analogous to the Zeno or Watch Pot effect.
242: \item
243: The interaction of $q$ with the environment $x_{\al}$ diminishes 
244: the quantum coherence. As a consequence, the system behaves more 
245: like a classical system than like a quantum system. Since tunneling is a 
246: pure quantum phenomenon, it should be suppressed as the system becomes 
247: more macroscopic/classical. We shall refer to this as decoherence, by which we mean
248: the process where a quantum system behaves more classically (i.e., less quantum)
249: via its interactions with the environment \cite{zeh}.
250: 
251: \end{itemize}
252: 
253: Here, we study this effect in quantum gravity, in the tunneling 
254: from nothing scenario. In this case, the cosmic scale factor $a$ 
255: plays the role of the system, while the metric fluctuations around $a$ (and any matter field modes)
256: play the role of $x_{\al}$, i.e., the environment. Figure 3 illustrates this situation. 
257: \begin{figure}
258: \begin{center}
259: \epsfig{file=ptbd_s4.eps, width=9cm}
260: \vspace{0.1in}
261: \caption{Typical tunneling involves metric and matter excitations as well. Loop effects may also be included, though they are not important for our discussions.}
262: \label{fig3}
263: \end{center}
264: \end{figure}
265: Since we measure only $a$, the metric fluctuations are integrated out 
266: in the path integral (or traced out in the density matrix).  
267: As expected, we shall show that their presence suppresses the tunneling
268: from nothing to a deSitter universe.
269: As expected, the decoherence effect is negligible for large cosmological constant 
270: (small size universe), but becomes
271: increasingly important as $\Lambda \to 0$, as 
272: it suppresses the tunneling probability when $a = 1/\sqrt{\Lambda/3}$ becomes macroscopic.
273: In contrast to a normal quantum system, the inclusion of the environment is of 
274: fundamental importance in quantum gravity, since the corrected Euclidean 
275: action is now bounded from below. 
276: 
277: The coupling of the metric fluctuations and scalar fields to $a(\tau)$ are more complicated than that of 
278: the above quantum system. Each metric perturbative mode behaves like a simple harmonic 
279: oscillator but with time-dependent (or $a$-dependent) mass and frequency. However, the real subtlety of the calculation comes in another way. If we treat the metric fluctuation modes as pure perturbations, 
280: we shall get nothing except the loop correction, a known result in Euclidean gravity. 
281: This is not hard to see. The Euclidean action in minisuperspace (that is, keeping only the cosmic scale factor) is given by
282: \ba
283: S_{E} = \frac{1}{2}\int d\tau \left( -a\dot{a}^2 - a + \lambda a^3 \right)
284: \ea
285: where a rescaling has rendered $\tau$, $a(\tau)$, the Hubble constant $H=1/\sqrt{\lambda}$ and $\lambda=2G\Lambda/9 \pi$ dimensionless.
286: For $S^4$, $Ha(\tau)= \sin (H \tau)$, where south pole (north pole) corresponds to $H \tau= 0$ ($\pi$). This path gives the Euclidean action (\ref{EactionG}). For a fluctuation mode $f$ that satisfies the classical equation, the Euclidean classical action can be written as a surface term
287: $S_E \simeq a^3 f {\dot f} {\large |}^{\pi/H}_0=0$, 
288: since $a(\tau)=0$ at the two poles. As a result, no decoherence term is generated. (Including an additional boundary term makes no difference.)
289: However, we find that the back-reaction is crucial for getting the correct answer. Instead of using the unperturbed $a(\tau)$ given above for the $S^4$ geometry, we leave it arbitrary during the tracing 
290: out of the fluctuation modes. 
291: In the path-integral formalism, one starts with the path integral that includes the scale factor $a$  as well as the perturbations around $a$. The following, for example, 
292: shows the inclusion of the metric tensor perturbations $t_n$ 
293: \ba
294: \label{path}
295: Z = \int D[a]\prod_{n}\int_{t_n^i}^{t_n^f}D[t_n]e^{-S_E} \exp\left( 
296:  -\frac{1}{2}\int_{-T/2}^{T/2} d\tau a^3 \sum_{n} [\dot{t_n}^2 + \omega_n(a)^2
297: t_n^2]  \right)
298: \ea
299: Tracing out the perturbations, we have 
300: \ba
301: \label{trace}
302: Tr[Z] = \prod_{n} \int dt_n^i \int dt_n^f \delta(t_n^i - t_n^f)~ Z
303: \ea
304: This results in a new term in the modified action  
305: \ba
306: S_{E,dC} \simeq \frac{1}{2}\int d\tau \left( -a\dot{a}^2 - a + \lambda a^3 + \frac{\nu}{\lambda^2 a}\right)
307: \ea
308: where the last term, coming from integrating out the perturbative modes, behaves like ordinary radiation. $\nu$ is a constant that measures the number of perturbative modes.  We then solve for $a(\tau)$ and obtain the corrected Euclidean action $S_{E,dC}$ in the saddle-point approximation.
309: It turns out that the $\nu$ term modifies the shape of the $S^4$ instanton to barrel-like, as shown in Figure 4. Since the effect is perturbative in nature, we expect the barrel to have the same topology as $S^4$. We see that, due to the back-reaction, $a(\tau)$ does not vanish at the two ends. However, the contribution of the end plates of the barrel to $S_{E,dC}$ happens to be zero.
310:  
311: After tracing over the metric fluctuations in this way, the rate 
312: of tunneling from nothing (i.e., no classical spacetime) to a 
313: deSitter universe (much like the inflationary universe) is now
314: given by $\Gamma \simeq \exp(F)= \exp(-S_{E,dC})$, where
315: \ba
316: \label{FS}
317: - F = S_{E,dC} = S_{E,0}+ D \simeq -\frac{3 \pi}{G\Lambda} + c \left(\frac{3 \pi}{G\Lambda}\right)^2
318: \ea
319: where $D$ is the decoherence term and $c$ depends on the cut-off.
320: In string theory, that cut-off is naturally provided by the string scale.
321: Note that $S_{E,dC}$ is now bounded from below. See Figure 5.
322: Here, $c$ in string theory also depends on the string spectrum, so 
323: $c$ should be calculated for each vacuum. 
324: %We shall call the second term the decoherence term.
325: We find that 
326:  \ba
327:  c = \frac{3}{2} n_{dof} \left( \frac{2 G}{3\pi l_{s}^2} \right)^2
328:  \ea
329: where $l_s$ is the string scale ($\alpha^{\prime} = (l_s/2 \pi)^2$) and 
330: $n_{dof}$ is the number of light degrees of freedom included in the environment.
331: For the pure gravity case, we have $n_{dof}=2$ for the two tensor modes.
332: For small $\Lambda$, below a critical value, the tunneling is actually totally suppressed.
333: \begin{figure}
334: \begin{center}
335: \epsfig{file=def_ball.eps, width=9cm}
336: \vspace{0.1in}
337: \caption{Modification of the $S^4$ instanton due to the presence of decoherence, 
338: i.e., the inclusion of the environment.}
339: \label{fig4}
340: \end{center}
341: \end{figure} 
342: 
343: This is quite understandable. A precise $S^4$ spherical geometry leads to the pure de Sitter space, which by definition excludes any radiation. If we do not allow back-reaction, then the system $a(\tau)$
344: actually cannot feel the presence of the environemnt. Allowing back-reaction, the quantum fluctuation
345: during the spontaneous creation of the universe generates some radiation (even though no 
346: radiation is introduced classically). They act as the environment.
347: The presence of this environment modifies the geometry in a way so that tunneling is to a universe with both a cosmological constant and some radiation, with a suppressed tunneling rate. 
348: The presence of the radiation generates the decoherence term. Note that, in integrating out the perturbative modes, the case with zero amplitude for the perturbative modes is also included.
349: As we see, the amount of radiation is proportional to $1/\Lambda^{2}$, so there is more radiation in a larger universe. 
350: 
351: What we consider fundamental is that the decoherence effect 
352: actually provides the Euclidean action $S_E$ of pure gravity with a lower bound.
353: Since the metric fluctuation contribution to $S_E$ cannot be 
354: turned off, they must be included in the evaluation of the tunneling rate. 
355: For usual quantum system, $S_0$ is bounded from below. So 
356: one may view this friction/environment/decoherence effect as 
357: a correction, albeit it may be very big. In quantum gravity, 
358: this effect resolves the unboundedness problem.
359: That the quantum fluctuation provides a natural source to cure the boundedness problem 
360: implies that quantum gravity is actually self-consistent in the macroscopic regime. 
361: 
362: \begin{figure}
363: \begin{center}
364: \epsfig{file=max_2.eps, width=9cm}
365: \vspace{0.1in}
366: \caption{$S_{E,dC}$ and $S_{E,0}$ as functions of the cosmological constant $\Lambda$.
367: We see that, with the inclusion of the decoherence effect, the Euclidean action for the
368: $S^4$ instanton  is now bounded from below. For large $\Lambda$, the semi-classical approximation breaks down. For small $\Lambda$, tunneling is actually completely suppressed.
369: }
370: \label{fig5}
371: \end{center}
372: \end{figure}
373: This effect renders 
374: the system more macroscopic and so less quantum. That is, as $a$ 
375: becomes large, its interaction with the metric fluctuations and matter fields should 
376: suppress the tunneling rate. Indeed, we show that the inclusion of 
377: the metric fluctuation decreases the tunneling rate, as expected.
378: Note that the decoherence term is not the usual perturbative 
379: quantum correction. For large $\Lambda$, where quantum correction 
380: is expected to be large, the decoherence term actually becomes negligible.
381: One can include the quantum corrections; however, they do not change 
382: the qualitative behavior for moderate values of $\Lambda$, i.e., $G \Lambda <<1$. 
383: 
384: Let us gain some idea of the magnitude of $F$. Following Eq.(\ref{FS}), we find the value of $\Lambda$ with maximum tunneling probability is
385: \ba
386: \Lambda_{max} = \frac{4 n_{dof} G}{\pi l_s^4}  \to F_{max}=\frac{3}{2n_{dof}} (2 \pi M_{Pl} l_s)^4
387: \ea
388: For the string scale a few orders of magnitude smaller than $M_{Pl}$, we can easily have $F \simeq 10^{10}$ or larger. On the other hand, the critical value of $\Lambda$ is $\Lambda_c=\Lambda_{max}/2$. For $\Lambda < \Lambda_c$, the barrel-shaped instanton is destroyed. At $\Lambda_c$,
389: $F(\Lambda_c)=0$, so the tunneling probability at $\Lambda$ close to $\Lambda_c$ is already negligibly small compared to that at $\Lambda_{max}$.
390: 
391: Having resolved the outstanding problem mentioned above, one 
392: may then apply this consistent tunneling approach to select 
393: a preferred vacuum in the cosmic landscape in string theory.
394: It is straightforward to generalize Eq.(\ref{FS}) to arbitrary spatial dimensions. In particular, for ten dimensional spacetime, we have
395: \ba
396: \label{FS10}
397: - F = S_{E,dC} \simeq S_{E,10}  + c \left(\frac{V_{10}}{l_s^{10}}\right)
398: \ea
399: where $S_{E,10}$ is the 10-D Euclidean action determined in mini-superspace and 
400: $V_{10}$ is the 10-dimensional volume of the instanton. 
401: Note that (see Ref\cite{Firouzjahi:2004mx}) $S_{E,10}$ reduces to $-{3 \pi}/{G\Lambda}$
402: for a vacuum where the extra 6 dimensions are compactified.
403: For each vacuum, $c$ depends on 
404: the spectrum. It may also depend on the compactification and dilaton moduli.
405: Knowing these properties of each vacuum allows one to calculate its 
406: tunneling probability from nothing. One may estimate the size of $c$ by comparing to the 4-D case:
407: \ba
408: c \simeq \frac{n_{dof} } {\pi M_{Pl}^2 l_s^2g_s^2}
409: \ea
410: with $8\pi G= 1/M_{Pl}^2$ and $g_s$ is the string coupling. Since the string scale is expected to be a few orders of magnitude below the Planck scale, we expect $c$ to be a small number.
411: In \cite{Firouzjahi:2004mx}, we consider the suppression
412: of tunnelling in the context of the spontaneous creation of the 
413: universe. The suppression happens due to the effect of both
414: the gravitational and matter perturbations on the bounce solution
415: to the Euclidean Einstein equations.  The general idea, loosely
416: speaking,  is to separate the ``universe'' into a ``system'' (the pure
417: gravitational bounce) and the ``environment'' (the perturbations). 
418: One is interested in measuring properties of the system and in order 
419: to do so one simply traces out the environmental degrees of freedom. The effect of 
420: the environment can be significant and such effects have been studied in 
421: various systems. In the previous work \cite{Firouzjahi:2004mx} 
422: we estimated this effect  on the tunnelling probability 
423: of the universe by considering the unperturbed deSitter space as the 
424: ``system'' and the ``environment'' consists of the metric perturbations and matter fields.
425: We see that the qualitative features are robust, though the details are somewhat more 
426: involved. (There we used $V_9$ instead of $V_{10}$ for Eq.(\ref{FS10})). 
427: For large $\Lambda$, the semi-classical approximation 
428: breaks down. For small $\Lambda$, additional decoherence effect may be important. 
429: (Here we have calculated only the leading order.)
430: Fortunately, the range of $\Lambda$ where the tunneling probability is largest seem to lie
431: in the region where the approximation is most reliable; and we are interested in vacua 
432: with large tunneling probabilities. Since the tunneling probability drops off rapidly as 
433: $\Lambda \to 0$ and $\Lambda \to \infty$, their precise values are not 
434: as important to us. With some luck, we may use the above formula to locate the preferred 
435: set of vacua in the cosmic string landscape. Note that $c$ depends on the spectrum at each
436: point in the landscape. 
437: 
438: As one can see in Figure 5, intermediate values of $\Lambda$, much like the inflationary universe that describes the history of our universe \cite{Guth:1981zm} seems to be preferred.
439: As in Ref\cite{Firouzjahi:2004mx}, we find phenomenologically that 10-dimensional deSitter-like 
440: vacua (with $F \simeq 10^9$ for instantons $S^{10}$, $S^5 \times S^5$, $S^4 \times S^3 \times S^3$) are not 
441: preferred while supersymmetric vacua in any dimension have essentially zero tunneling probability. Also, tunneling to vacua very much like our today's universe (with a very small dark energy) seems to be severely suppressed. Tunneling to a universe with quantum foam seems not preferred either. Among the known vacua, the preferred ones are the 4-D brane inflation \cite{Dvali:1998pa} as realized in a realistic string model \cite{Kachru:2003sx}, very much like the inflationary universe that our universe has gone through (with $F \sim 10^{16}$). Although the details of the decoherence term obtained here is somewhat different from that used in Ref\cite{Firouzjahi:2004mx}, we see that the qualitative features summarized there remain true.
442: 
443: Decoherence and related issues in quantum/Euclidean gravity have been studied earlier 
444: \cite{Halliwell:1985eu,Kiefer:1987ft,Kiefer:1989ud,Halliwell:1989vw,Kiefer:1992cn}, where they are
445: mostly concerned with the evolution of the inflationary universe. Here, we study decoherence in the 
446: quantum tunneling in gravity. To be self-contained, we shall review some of the relevant formalisms developed there.
447: 
448: In Sec. 2, we review the tunneling to closed deSitter space and perturbative modes around 
449: the cosmic scale factor $a(\tau)$.
450: In Sec. 3, we evaluate the contribution due to these modes to the effective Euclidean action. Here we find that no decoherence term is generated in the $S^4$ background. To see why back-reaction is important, the effect is calculated in the  background of a ``squashed'' $S^4$ geometry. The reader may choose to skip this section. 
451: In Sec. 4, we study the modified bounce including back reaction. We derive the effective Euclidean action after integrating out the metric perturbation with an arbitrary background $a(\tau)$.
452: In Sec. 5, We find the modified bounce solution from the effective Euclidean action. This allows 
453: us to obtain the improved wavefunction and tunneling probability from nothing. 
454: In Sec. 6, we show that the new term in the improved Euclidean action behaves like ordinary radiation.
455: We see also how the new term appears in the Wheeler-DeWitt equation and how it influences the tunneling amplitude. 
456: In Sec. 7, we discuss the implication of the main result and its connection to the proposal of 
457: Ref\cite{Firouzjahi:2004mx}.
458: Sec. 8 contains some discussions and Sec. 9 contains a summary and further remarks. Some of the details are relegated to appendices.
459: For the sake of completeness, a number of known results are reviewed extensively.
460: 
461: \section{Setup}
462: 
463: \subsection{Notations}
464: 
465: We shall render various quantities dimensionless for the ease of calculation.
466: The conventions we follow are those of \cite{Hartle:1983ai}. The 
467: Euclidean action is defined as
468: \ba
469: S_{E} =-\frac{1}{16 \pi G}\int d^4 x \sqrt{-g} \left( R - 2\Lambda\right)
470: \ea
471: The Euclidean metric is given by
472: \ba
473: ds^2 = \sigma^2 \left( d\tau^2 + a(\tau)^2 d\Omega_{3}^2 \right)
474: \ea
475: where is $\sigma^2 = {2G}/{3 \pi}$. With this metric ansatz, the 
476: action becomes
477: \ba
478: S_{E} = \frac{1}{2}\int d\tau \left( -a\dot{a}^2 - a + \lambda a^3 \right)
479: \ea
480: where $\lambda = {\sigma^2 \Lambda}/{3} = {2G \Lambda}/{9 \pi}$. So
481: $\lambda$, $\tau$, and $a$ are all dimensionless.
482: 
483: 
484: %\section{The deSitter Spacetime and its Perturbations}
485: 
486: \subsection{The deSitter Space}
487: 
488: In this section we give a summary of the result of \cite{Halliwell:1985eu}.
489: Consider a compact three-surface $S^3$ which divides the four-manifold M into
490: two parts. One can introduce the coordinates $x^i$ ($i = 1,2,3$) and a coordinate $t$ 
491: such that $S^3$ is the surface at $t = 0$. The metric takes the form
492: \ba
493: ds^2 = -(N^2 - N_i N^i)dt^2 + 2N_idx^idt + h_{ij}dx^idx^j.
494: \ea
495: where $N$ and $N_i$ are the lapse function and the shift vector, respectively.
496: The action is given by
497: \ba
498: S = \int \left( L_g + L_m \right) d^4x
499: \ea
500: where
501: \ba
502: L_g = \frac{M_{P}^{2}}{16\pi}N \left( G^{ijkl}K_{ij}K_{kl}
503: + \sqrt{h} R  \right)
504: \ea
505: where $R$ is the Ricci scalar of the three-surface and $K_{ij}$ is
506: the second fundamental form given by
507: \ba
508: K_{ij} = \frac{1}{2N}\left( -\frac{\partial h_{ij}}{\partial t}
509: + 2N_{(i|j)}  \right)
510: \ea
511: In the above expression ``$|$'' denotes the covariant derivative.
512: $G^{ijkl}$ is called the metric of the ``superspace'' and is given by
513: \ba
514: G^{ijkl} = \frac{1}{2}\sqrt{h}\left(h^{ik}h^{jl}+ h^{il}h^{jk}
515: - 2h^{ij}h^{kl} \right)
516: \ea
517: In the case of a massive scalar field, the matter Lagrangian $L_m$
518: is given by
519: \ba
520: L_m = \frac{1}{2}N\sqrt{h}\left( N^{-2}\left(\frac{\partial \Phi}
521: {\partial t} \right)^2 -2\frac{N^i}{N^2}\frac{\partial \Phi}
522: {\partial t} \frac{\partial \Phi}{\partial x^i} 
523: \right . \\ \nonumber \left. - \left( h^{ij} 
524: -\frac{N^i N^j}{N^2}\right)\frac{\partial \Phi}
525: {\partial x^i}\frac{\partial \Phi}{\partial x^j} - m^2 
526: \Phi^2  \right)
527: \ea
528: In the Hamiltonian treatment of general relativity one treats $h_{ij}$
529: and $\Phi$ as the canonical coordinates. The canonically conjugate
530: momenta are
531: \ba
532: \pi^{ij} = \frac{\partial L_g}{\partial\dot{h_{ij}}}
533: = -\frac{\sqrt{h}M_P^2}{16\pi}\left(K^{ij} - h^{ij} K \right)\\
534: \nonumber
535: \pi_{\Phi} = \frac{\partial L_m}{\partial \dot{\Phi}}
536: = N^{-1}\sqrt{h}\left(\dot{\Phi}-N^i \frac{\partial \Phi}{\partial x^i}   
537: \right)
538: \ea
539: The Hamiltonian is
540: \ba
541: H = \int d^3x \left(\pi^{ij}\dot{h_{ij}} + \pi_{\Phi}
542: \dot{\Phi}-L_g -L_m \right) \\ \nonumber
543: = \int d^3x \left(NH_0 + N_i H^i \right)
544: \ea
545: where
546: \ba
547: H_0 = 16 \pi M_P^{-2}G_{ijkl}\pi^{ij}\pi^{kl}  - \frac{M_P^{2}}{16\pi}
548: \sqrt{h}R + & \\ \nonumber\frac{1}{2}\sqrt{h}\left(\frac{\pi_{\Phi}^2}{h}
549: + h^{ij} \frac{\partial \Phi}{\partial x^i}
550: \frac{\partial \Phi}{\partial x^j} 
551:  + m^2 \Phi^2 \right) 
552: \ea
553: \ba
554: H^i = -2\pi^{ij}_{|j} + h^{ij}\frac{\partial \Phi}{\partial x^j}
555: \pi_{\Phi} &&
556: \ea
557: and
558: \ba
559: G_{ijkl} = \frac{1}{2}h^{-1/2}\left(h_{ik}h_{jl}+h_{il}h_{jk}-
560: h_{ij}h_{kl} \right)
561: \ea
562: The quantities $N$ and $N_i$ are regarded as Lagrange multipliers.
563: Thus the solution obeys the momentum constraint
564: \ba
565: H^i = 0
566: \ea
567: and the Hamiltonian constraint
568: \ba
569: H_{0} = 0.
570: \ea
571: 
572: \subsection{The Perturbations}
573: 
574: Now we study the perturbations around the deSitter spacetime. The perturbed
575: deSitter has a three-metric $h_{ij}$ of the form
576: \ba
577: h_{ij} = a^2 \left( \Omega_{ij}+\epsilon_{ij}\right)
578: \ea
579: where $\Omega_{ij}$ is the metric on the unit three-sphere and 
580: $\epsilon_{ij}$ is a perturbation on this metric and can be expanded
581: in harmonics:
582: \ba
583: \label{harmo}
584: \epsilon_{ij} = \sum_{n,l,m} \left[ \sqrt{6}a_{nlm}\frac{1}{3}
585: \Omega_{ij} Q^n_{lm} + \sqrt{6}b_{nlm}(P_{ij})^n_{lm} +\right . &  \nonumber \\
586: \left .
587: \sqrt{2}c^0_{nlm}(S^0_{ij})^n_{lm} + \sqrt{2}c^e_{nlm}(S^e_{ij})
588: ^n_{lm} + 2t^0_{nlm}(G^0_{ij})^n_{lm}
589: + 2t^e_{nlm}(G^e_{ij})^n_{lm}
590: \right]
591: \ea
592: where the coefficients $a_{nlm}$, $b_{nlm}$, $c^0_{nlm}$, $c^e_{nlm}$,
593: $t^0_{nlm}$ and $t^e_{nlm}$ are functions of time but not the three 
594: space coordinates. 
595: The $Q(x^i)$ are the scalar harmonics on the three-sphere. The
596: $P_{ij}(x^i)$ are given by
597: \ba
598: P_{ij} = \frac{1}{n^2 -1}Q_{|ij}+ \frac{1}{3}\Omega_{ij}Q
599: \ea
600: where we have suppressed the $n,l,m$ indices. The $S_{ij}$ are
601: given by
602: \ba
603: S_{ij} = S_{i|j} + S_{j|i}
604: \ea
605: where $S_{i}$ are the transverse vector harmonics. The $G_{ij}$
606: are the transverse traceless tensor harmonics. Further details 
607: can be found in \cite{Halliwell:1985eu,Gerlach:1978gy}.
608: The lapse, shift, and the scalar field $\Phi(x^i,t)$ can be
609: expanded in terms of harmonics as
610: \ba
611: N = N_0 \left[1 + \frac{1}{\sqrt{6}}\sum_{n,l,m}g_{nlm}Q^n_{lm} \right]
612: \\ \nonumber
613: N_i = a(t)\sum_{n,l,m}\left[\frac{1}{\sqrt{6}}k_{nlm}(P_i)^n_{lm}
614: + \sqrt{2}j_{nlm}(S_i)^n_{lm} \right] \\ \nonumber
615: \Phi = \sigma^{-1}\left[\frac{1}{\sqrt{2}\pi}\phi(t)+ \sum_{n,l,m}f_{nlm}Q^n_{lm} \right]
616: \ea
617: where $P_i = \frac{1}{n^2 -1}Q_{|i}$. The perturbed action is
618: now given by
619: \ba
620: \label{act}
621: S = S_{0}(a, \phi, N_0)+ \sum_{n} S_n
622: \ea
623:  where we have denoted the labels $n$, $l$, $m$, $o$, and $e$ by the single
624: label $n$. $S_0$ is the action of the unperturbed deSitter,
625: \ba
626: S_{0} = -\frac{1}{2}\int dt N_0 a^3\left(\frac{\dot{a}^2}{N_o^2a^2} 
627: - \frac{1}{a^2}- \frac{\dot{\phi}^2}{N_0^2}+ m^2\phi^2 + \lambda
628:  \right) 
629: \ea
630: $S_n$ is quadratic in perturbations and is given by
631: \ba
632: S_n = \int dt (L^n_g + L^n_m)
633: \ea
634: where $L^n_g$  and $L^n_m$ are the $n$th mode gravitational and
635: the matter Lagrangians, respectively. As we are interested only
636: in the gravitational tensor and the scalar field perturbations
637: in this paper, we only display those terms here. In the absence of sources, the 
638: other modes can be gauged away. For further
639: details and the intricacies of all perturbations we refer the
640: reader to \cite{Halliwell:1985eu}. 
641: 
642: The tensor perturbations $t_n$ have the Euclidean action
643: \ba
644: S_n^E = \frac{1}{2}\int t_n \hat{D} t_n +  ~boundary ~term
645: \ea
646: where
647: \ba
648: \hat{D} = \left(- \frac{d}{d\tau}\left[a^3\frac{d}{d\tau} \right]
649: + a(n^2 - 1)  \right)
650: \ea
651: where the background satisfied the classical equation of motion.
652: The action is extremized when $t_n$ satisfies the equation
653: $\hat{D} t_n = 0$. Setting $N_0=1$, this is just the equation
654: \ba
655: \label{eom_ten}
656: \frac{d}{dt}\left[a^3 {\dot{t_n}} + (n^2 -1)a t_n\right] = 0
657: \ea
658: For $t_n$ that satisfies the equation of motion, the action
659: is just the boundary term
660: \ba
661: \label{E_cl}
662: S_n^{E(cl)} = \frac{1}{2}a^3 \left(t_n\dot{t_n}+4\frac{\dot{a}}{a}
663: t_n^2  \right)
664: \ea 
665: The path integral over $t_n$ will be
666: \ba
667: \label{tensor}
668: \int d[t_n]\exp (-S_n^{E(cl)})= (\det \hat{D})^{-1/2}\exp (-S_n^{E(cl)})
669: \ea
670: 
671: Scalar fields can be treated in a similar fashion. The scalar field perturbation Lagrangian is given by
672: \ba
673: \frac{1}{2}N_0 a^3 \left[\frac{1}{N_0^2}\dot{f_n}^2 -m^2 f_n^2 -\frac{(n^2 -1)}{a^2}f_n^2 \right]
674: \ea
675: where we work in an appropriate gauge choice. 
676: %Also we have assumed that the scalar field kinetic term is negligible. 
677: Setting $N_0=1$, the equation of motion for $f_n$ is
678: \ba
679:  \frac{d}{dt}\left(a^3 {\dot{f_n}}\right) + \left[m^2 a^3 + (n^2 -1 )a \right]f_n = 0
680: \ea
681: One can evaluate the scalar field path integral and the results are 
682: similar to that of the tensor perturbative modes. So in this paper
683: we deal only with the gravitational tensor modes.
684: 
685: 
686: \section{Perturbative Correction to the Bounce : No Back Reaction}
687: 
688: Before doing the calculation that includes the backreaction due to the
689: perturbative modes, we first perform a simple calculation to illustrate the key issues we are facing.
690: The result of this section is meant to be a warm up and indicates the possibility
691: of a correction to the wavefunction. We consider a $S^4$ solution with both its polar regions flattened by a small parameter $\delta$ and show that this squashed geometry allows for extra perturbative modes that can have significant effect on the calculation of the wavefunction. When $\delta = 0$, these extra perturbative modes vanish and what remains is the well known one-loop correction to the $S^4$.  For a proper treatment the reader
692: may go directly to the next section and onward.\\
693: 
694:  The Euclidean equation of motion for the $n$th tensor perturbation mode
695: on a $S^4$ background follows from Eq.(\ref{eom_ten})
696: \ba
697: \label{t_n}
698: \ddot{t_n}+ 3\frac{\dot{a_o}}{a_o}\dot{t_n} - 
699: \left( \frac{(n^2 -1)}{a_o^2} \right)t_n  = 0
700: \ea
701: where dot denotes differentiation with respect to the $\tau$ variable and
702: $a_o(\tau) = \lambda^{-1/2} \sin(\sqrt{\lambda}\tau)$.
703: This can be converted to a more familiar form by the substitution
704: $F_n = \sqrt{\lambda} a_o(\tau) t_n$.
705: In terms of $F_n$ and $x = \cos(\lambda \tau)$ the equation of motion becomes
706: \ba
707: (1 - x^2)\frac{d^2 F_n}{dx^2} - 2x\frac{dF_n}{dx} +
708: \left( 2 - \frac{n^2}{(1 - x^2)} \right) F_n = 0 
709: \ea
710: which is just an associate Legendre equation of degree one
711: and order $n$. This has two linearly
712: independent solutions, $P^{-n}_{1}(x)$ and $Q^{n}_{1}(x)$.
713: However, because $a_o=0$ at $x=\pm 1$, we have
714: \ba
715: S_{E}^{n} & = & \frac{1}{2} \left[a(\tau)^3 t_n \frac{dt_n}{d\tau}  
716: \right]_{x = 1}^{x = -1} =0
717: \ea
718: 
719: We shall see in the following sections that once the perturbative modes
720: are properly accounted for, $S^4$ changes to a ``barrel''. Anticipating
721: this result we consider the $S^4$ instanton that is flattened slightly
722: at the two poles. This ``squashed'' $S^4$
723: is given by $a_o(\tau) = \frac{1}{\sqrt{\lambda}}\sin(\lambda \tau)$,
724: with the restricted range $ 1 - \delta \geq x \geq -1 + \delta$. There
725: is a good reason for considering the squashed geometry. The perturbative
726: modes can be strong enough to change the geometry of $S^4$ (and as we shall
727: see, they will). So fixing the geometry to $S^4$ before doing the
728: perturbative analysis is too restrictive. The perturbative modes 
729: on $S^4$  have to vanish at the two poles if they are to respect the
730: background geometry. Their effect has been studied in \cite{Barvinsky:1992dz
731: ,Gibbons:1978ji} and apart from contributing to one-loop effect 
732: they do not lead to any tunneling suppression.
733:  However, to get the decoherence effect  one must
734: include modes that have nonvanishing values at the two poles. And
735: we shall see that the squashed geometry allows for modes that
736: can potentially lead to decoherence. This expectation will be confirmed in
737:  the later sections. \\
738: 
739:      As explained in the introduction (Eq.(\ref{path},\ref{trace})), to 
740: trace out a given  perturbative mode
741: (say, $t_n$) one must perform a path integral over that mode with
742: the initial and final amplitudes the same (say, $t_n^i = t_n^f$), and
743: then integrate over all possible values of $t_n^i$. This is just
744: the trace operation in the path integral formalism. To do this we
745: first find the action for $t_n$ (details can be found in the appendix)
746: 
747: \ba
748: S_{E}^{n} & = & \frac{1}{2} \left[a(\tau)^3 t_n \frac{dt_n}{d\tau}  
749: \right]_{x = (1 - \delta)}^{x = (-1 + \delta)} \\ \nonumber
750: & = & \delta(2 -\delta) \left[ (1 - \delta) + \delta(2 -\delta)
751: \frac{\partial_{x}Q^{n}_1(1 -\delta)}{Q^{n}_1(1 -\delta)}\right] (t_n^i)^2
752: \ea
753: Next we must find the prefactor to the path integral for $t_n$.
754: This can be found as explained in the Appendix B. The prefactor 
755: is given by
756: \ba
757: \frac{1}{\sqrt{4\pi}} \frac{\sqrt{Q^n_1(1 -\delta) P^n_1(1 - \delta) 
758: (2\delta- \delta^2)}}{2^n} \sqrt{\frac{\Gamma\left( \frac{2-n}{2}
759: \right) \Gamma\left( \frac{3-n}{2}\right)}{\Gamma\left( 
760: \frac{2+ n}{2}\right) \Gamma\left( \frac{3+n}{2}\right)} }
761: \ea
762: Integrating over all initial states $t_n^i$ (tracing out the $nth$ mode)
763: gives
764: \ba
765: \int dt_n^i K \left( t_n^i, x = -1 + \delta ;t_n^i, x = 1 -\delta \right)  
766: = \alpha (n) f(\delta) 2^{-n}
767: \ea
768: where $\alpha(n)$ contains factors of $n$ and $f(\delta)$ is a function 
769: of $\delta$, with $f(0)=0$. $K \left( t_n^i,-1 + \delta ;t_n^i,1 -
770: \delta \right)$ is the path integral over $t_n$ mode with the boundary 
771: conditions
772: $t_n(x = -1 + \delta) = t_n(x = 1 - \delta) = t_n^i$.
773: Taking care of all the modes by tracing over all of them (within the
774: lower and upper cut-offs), we get
775: \ba
776: \prod_{n}\int dt_n^i  K \left( t_n^i, -1 + \delta ;t_n^i, 1 -\delta \right)
777: = \prod_{n}
778: \alpha (n) f(\delta) 2^{-n} \\ \nonumber
779: = A(n, \delta)e^{-N^4/2 \ln(2)}
780: \ea
781: where $N$ counts the modes. As explained later,
782: %in Eq.(\ref{N}, \ref{nu})
783: \ba
784: N = \left( \frac{H^{-1}}{l_s}\right) = 2^{1/4}\frac{\nu^{1/4}}{\sqrt{\lambda}}
785: \ea
786: where $\nu = \frac{2}{9\pi^2}\frac{G^2}{l_s^4}$. The 
787: wavefunction then corresponds to the path integral over the scale factor
788: $a$ and the perturbative modes $t_n$ with the full action as given
789: in Eq.(\ref{act}). We trace over the perturbative modes and get
790: \ba
791: \Psi \simeq e^{\left(\frac{1}{3\lambda} - D \right)}
792: \ea
793: where $D$ is the decoherence term leading to tunneling suppression
794: \ba
795: D \simeq \frac{\nu}{\lambda^2}
796: \ea
797: so the above result leads 
798: to an order O(${1}/{\lambda^2})$ suppression to the Hartle-Hawking
799: wavefunction. This naive (naive because, as it turns out, the backreaction
800: is important and also leads to an O(${1}/{\lambda^2})$ contribution)
801: expectation is indeed vindicated in our calculation of the modified bounce.
802: The more careful calculation changes the coefficient $\ln(2)$, but maintains
803: the inverse square dependence on $\lambda$.
804: 
805: 
806: \section{The Modified Bounce : Including Backreaction}
807: 
808: The bounce solution $S^4$ is a solution to the Euclidean Einstein
809: equation which is obtained as the Euler-Lagrange equation from
810: the Euclidean action $S^E_{0}[a]$. 
811: \ba
812: \label{euc_eins}
813: -2\frac{\ddot{a}}{a} - \frac{\dot{a}^2}{a^2}+\frac{1}{a^2}
814: = 3 \lambda
815: \ea
816: The solution is given by $a(\tau) = 
817: \sqrt{\frac{1}{\lambda}}\cos(\sqrt{\lambda}\tau)$. This is 
818: the bounce in the absence of any perturbation. Including
819: the perturbations, treating them as the environment, and tracing
820: them out (as explained in Eq.(\ref{path},\ref{trace})) will lead to 
821: a modified equation of motion instead
822: of Eq.(\ref{euc_eins}). In this section we derive this modified
823: bounce equation.
824: 
825: Let us consider the effect of the metric perturbations. To find 
826: the modified bounce equation we have to carry out the
827: path integral over the perturbation modes $t_{n}$ is Eq.
828: (\ref{tensor}). Let us write down the path integral for a single
829: tensor perturbation mode
830: \ba
831: \label{PI}
832: \int D[t_{n}(\tau)]\exp \left( -\frac{1}{2}\int_{-T/2}^{T/2}
833: d\tau a^3 [ \dot{t_{n}}^2 + \frac{(n^2 - 1)}{a^2} 
834: t_{n}^2 ] \right)
835: \ea
836: This is a path integral for an oscillator with a varying
837: mass as well as frequency. We can simplify this to a path
838: integral of an oscillator with constant mass and variable
839: frequency (given by $\omega_n (a(u)) = \sqrt{(n^2 -1)}a(u)^2$) 
840: using a new variable $u$
841: \ba
842: du = \frac{d\tau}{a(\tau)^3}
843: \ea
844: The Euclidean action now becomes
845: \ba
846: \label{action}
847: S_{E} = \frac{1}{2}\int_{u_i}^{u_f}du \left( t_{n}'^2 +
848:  (n^2 - 1)a^4t_{n}^2\right) \\ \nonumber
849: = \frac{1}{2}\int_{u_i}^{u_f} du \left( -tt'' + \omega_n^2 t^2
850: \right) + \frac{1}{2} (t{t}'){\Large\bf{ |}}^{u_f}_{u_i}
851: %\int_{u_i}^{u_f} du \frac{d}{du}\left(tt'\right)
852: \ea
853: To keep notation uncluttered, we drop the subscript $n$ for
854: the time being. Let $t = t_{cl} + \hat{t}$, where
855: $t_{cl}$ is a solution to the classical equation of motion
856: for the above action
857: \ba
858: \label{eom}
859: t_{cl}'' - \omega_{n}(a)^2 t_{cl} = 0
860: \ea
861: with $t_{cl}(u_i)=t(u_i)$ and $t_{cl}(u_f)=t(u_f)$. 
862: Here $\hat{t}$ denotes fluctuations about
863: the classical solution with $\hat{t}(u_i)=\hat{t}(u_f)=0$. That is,
864: $(t {t}'){\Large |}^{u_f}_{u_i}= (t_{cl}\hat{t}'){\Large |}^{u_f}_{u_i} + (t_{cl}{t_{cl}}'){\Large |}^{u_f}_{u_i}$.
865: Apriori, the contributions to the path integral 
866: will come from $t_{cl}$ and $\hat t$.
867: Here, the fluctuations $\hat{t}$ will lead to a prefactor. We shall keep
868: track of this prefactor as it will have important contribution
869: to the modified action. Substituting $t = t_{cl} + \hat{t}$ in 
870:  Eq.(\ref{action}), we obtain
871: \ba
872: \label{actionI}
873: S_E = \frac{1}{2}\int_{u_i}^{u_f} du \left( -\hat{t}\hat{t}''
874: + \omega_n^2 \hat{t}^2 \right) + (t_{cl}{t_{cl}}'){\Large\bf{ |}}^{u_f}_{u_i}
875: %\frac{1}{2}\int_{u_i}^{u_f} du \frac{d}{du}(t_{cl}t_{cl}')
876: \ea
877: The second term in the above equation is simply $S_E(t_{cl})$.
878: The path integral in Eq.(\ref{PI}) is, therefore, given by
879: \ba
880: \label{PII}
881: e^{-S_E(t_{cl})} P[{\hat{t}}]=e^{-S_E(t_{cl})} \int^0_0 D[\hat{t}(u)] \exp \left( -\frac{1}{2}\int_{u_i}^{u_f} 
882: du \left( -\hat{t}\hat{t}''+ \omega_n^2 \hat{t}^2 \right)\right)
883: \ea
884: Notice that the prefactor $P[{\hat{t}}]$ is independent of $t(u_i)$ or $t(u_f)$.
885: So the reduced path integral Eq.(\ref{trace}) becomes
886: \ba
887: \label{trace2}
888: Tr[Z] =  \int D[a] \prod_{n} P[{\hat{t_n}}] \int dt_n^i \int dt_n^f \delta(t_n^i - t_n^f) e^{-S_E(t_{n,cl})}
889: \ea
890: We shall first evaluate the integral over $t_n^i=t_n^f$. In Appenxix C we evaluate the prefactor 
891: $P[{\hat{t_n}}]$. It is straightforward to include a scalar field.
892: 
893: Although finding a general solution to Eq.(\ref{eom}) is
894: in general impossible, we note that it has the
895: same form as the Schrodinger equation for a particle in
896: a potential $\omega_n (a(u))$. So for a slowly varying
897: potential we can use the WKB method for finding the solution
898: to Eq.(\ref{eom}). By slowly varying one means that the
899: following condition is satisfied
900: \ba
901: \frac{d\omega_{n}}{d\tau} << \omega_{n}^2
902: \ea
903: This is satisfed by the higher modes $n >> 1$ and we 
904: shall see that these modes are the ones that contribute  
905: to the suppression of quantum tunneling. 
906: (One may be concerned that $ \omega_{n} \to 0$ as $a \to 0$.
907: As we shall see, in contrast to the unperturbed bounce, $a$ actually stays 
908: finite in the modified bounce. In this sense, the WKB approximation is 
909: reasonable.)
910: The two independent WKB solutions to Eq.(\ref{eom}) are
911: \ba
912: \label{class_soln}
913: t_{cl}^{\pm} = \frac{1}{\sqrt{\omega_n}}\exp \left( \pm 
914: \int^u du' \omega_n(u')\right)
915: \ea
916: The general solution will be a linear combination of the
917: independent solutions
918: \ba
919: \label{gensol}
920: t_{cl}(u) = At_{cl}^{+}(u) + Bt_{cl}^{-}(u)
921: \ea
922: The values of $A$ and $B$ will depend on the boundary
923: conditions. If Eq.(\ref{gensol}) is to satisfy the 
924: following boundary conditions
925: \ba
926: t_{cl}(u_i) = t^i &  t_{cl}(u_f) = t^f
927: \ea
928: then we have the following values of $A$ and $B$
929: \ba
930: \label{AB}
931: A = \frac{t^f \sqrt{\omega_n(u_f)} - t^i 
932: \sqrt{\omega_n(u_i)}\exp(- D_n)}{(\exp(D_n) - \exp(-D_n))} 
933: \\ \nonumber
934: B = -\frac{t^f \sqrt{\omega_n(u_f)} - t^i 
935: \sqrt{\omega_n(u_i)}\exp(D_n)}{(\exp(D_n) - \exp(-D_n))}
936: \ea
937: where $D_{n}$ is given by
938: \ba
939: \label{D_neq}
940: D_n = \int_{u_i}^{u_f}du ~ \omega_{n}(a(u)) = 
941: \int_{-T/2}^{T/2}d\tau~ \frac{\omega_{n}(a(\tau))}{a(\tau)^3}
942: \ea
943: Next, substituting the solution Eq.(\ref{gensol}), using 
944: Eq.(\ref{AB}), in Eq.(\ref{actionI}), we get
945: \ba
946: \label{S_E}
947: & S_{E}(t_{cl}) = \frac{1}{2} \left( t_{cl}(u_f)\frac{dt_{cl}(u_f)}{du} - 
948: t_{cl}(u_i)\frac{dt_{cl}(u_i)}{du} \right)   \\ \nonumber
949: & = \frac{1}{2(\exp(D_n) - \exp(-D_n))} \left [\left( (t^f)^2 
950: \omega_n(u_f) +(t^i)^2 \omega_n(u_i) \right) \left( \exp(D_n) 
951: + \exp(-D_n)\right) \right. \\ \nonumber &  \left. - 4t^i t^f 
952: \sqrt{\omega_n(u_i)\omega_n(u_f)}
953: \right]
954: \ea 
955: This gives the classical contribution Eq.(\ref{PII}). Note that
956: in order to perform the trace over this perturbation mode, we
957: will be setting $t^i = t^f$, and $\omega_n(u_i) = \omega_n(u_f)$.
958: The path integral over $\hat{t}$ leads to the prefactor. Its evaluation
959: for a time-dependent frequency is explained in Appendix C (Eq.(\ref{go})).  
960: (Equivalently, one may use the Pauli-van Vleck-Morette formula  
961: studied by Barvinsky \cite{Barvinsky:1993nf}). 
962: The prefactor (path integral over $\hat{t}$) gives
963: \ba
964: \label{pre_fac}
965: \int D[\hat{t}(u)] \exp \left( \frac{1}{2}\int_{u_i}^{u_f} 
966: du \left( -\hat{t}\hat{t}'' + \omega_n^2 \hat{t}^2 \right)\right)
967: =  \frac{\sqrt{\omega_n(u_f)}}{\sqrt{\left(\exp(D_n) - \exp(-D_n)\right)}}
968: %\frac{1}{\sqrt{(\exp(D_n) - \exp(-D_n))}}
969: \ea
970: Now the trace over the perturbative modes can be performed by
971: setting $t^i = t^f$ and integrating over the amplitudes $t^i$
972: \ba
973: & \int dt^f \int dt^i \delta(t^i -t^f)\int_{t^i}^{t^f}D[t(u)] \exp 
974: \left( -S_{E}\right) \\ \nonumber
975: = &  \frac{\sqrt{\omega_n(u_f)}}{\sqrt{(\exp(D_n) - \exp(-D_n))}} 
976: \int dt^f \exp\left( -S_{E} [t_{cl}] \right)
977: \\ \nonumber
978: \simeq  & \frac{1}{\sqrt{\left(\exp(D_n) - \exp(-D_n)\right)}}
979: \ea
980: where we used $\int dt^f \exp\left( -S_{E} [t_{cl}]\right) \simeq 1/\sqrt{
981: \omega_n(u_f)}$. Also, note that $D_n >> 1$ for large values of $n$, and 
982: $\left(\exp(D_n) -  \exp(-D_n)\right)^{-1/2} \simeq \exp (-D_n/2)$ . 
983: Finally, one has to  perform this tracing-over over all $n$ modes. 
984: This leads to the following contribution
985: \ba
986: \prod_{n} \frac{1}{\sqrt{(\exp(D_n) - \exp(-D_n))}}\simeq \exp (- \sum\frac{D_n}{2} )
987: \ea
988:  From the expansion in Eq.(\ref{harmo})
989: it is clear that one has to count the indices $n, l, m$  (corresponding 
990: to the spherical harmonics on $S^3$). For a given $n$, $l$ can take 
991: $n-2$ values ($l = 2, 3, ..., n-1$), and for a given $l$, $m$ can take
992: $2l + 1$ values ($m = -l, -l + 1, ...., l - 1, l$). For a given $n$ this
993: introduces a degeneracy of $f(n) \simeq 2 n^2$. The factor of $2$ comes \
994: due to the two tensor modes
995: $t^0_{nlm}$ and $t^e_{nlm}$ in the expansion in Eq.(\ref{harmo}).
996: To proceed further, we need to define the cut-offs. A natural short wavelength 
997: cut-off is the string scale, $l_{s}$. The long wavelength cut-off
998: is the inverse Hubble length.
999: %The number of modes between these two cutoffs is given by $N$
1000: This gives a cut-off for $n$
1001: \ba
1002: \label{N}
1003: n_{max} = N = \left( \frac{H^{-1}}{l_s}\right)=\frac{1}{l_s}\sqrt{\frac{3}{\Lambda}}
1004: \ea
1005: \ba
1006: D \equiv \sum \frac{D_n}{2}  = \frac{1}{2} \sum^N_{n,l,m} \int_{-T/2}^{T/2}
1007: d\tau \frac{\omega_n(a)}{a^3} 
1008: \ea 
1009: where we have used Eq.(\ref{D_neq}).
1010: Note that 
1011: $\omega_n(a)^2 = (n^2 - 1)a^4$. For higher modes, therefore,
1012: $\omega_n \simeq na^2$. Thus,
1013: \ba
1014: D = \frac{1}{2}\int_{-T/2}^{T/2}\frac{d\tau}{a} \sum^N_{n,l,m} n \\ \nonumber
1015: = \frac{N^4}{8}\int_{-T/2}^{T/2}\frac{d\tau}{a}=\frac{1}{2}\int_{-T/2}^{T/2}{d\tau}\frac{\nu}{\lambda^2a}
1016: \ea
1017: We can also easily generalize to arbitrary number of degrees of freedom, with
1018: \ba
1019: \frac{\nu}{\lambda^2}= \frac{n_{dof} N^4}{4} = \frac{n_{dof}}{4(Hl_s)^4}
1020: \ea
1021: For pure gravity, we have $n_{dof}=2$.
1022: The modifed Euclidean action is, therefore, given by
1023: \ba
1024: \label{bounce}
1025: S_{E,dC}= S_{E,0} + D[a] = \frac{1}{2}\int d\tau
1026: \left( -a\dot{a}^2 - a + \lambda a^3 + \frac{\nu}{\lambda^2 a} \right)
1027: \ea
1028: To evaluate $S_{E,dC}$ via the saddle-point approximation, we have to find the classical path, or the bounce solution, of this effective action.
1029: 
1030: \section{The Bounce Solution}
1031: 
1032: The equation of motion for the Euclidean action,
1033: Eq.(\ref{bounce}), is
1034: \ba
1035: \label{modbounce}
1036: -2\frac{\ddot{a}}{a} - \frac{\dot{a}^2}{a^2}+\frac{1}{a^2}
1037: = 3\lambda - \frac{\nu}{\lambda^2 a^4} 
1038: \ea
1039: This is just the Euclidean version of the Einstein equation with
1040: both a cosmological constant and radiation. As is well known, this 
1041: equation allows for a variety of solutions (M1, M2, A1, A2, E, O1)
1042: \cite{Harrison:1967ab}. One has to be careful in deciding which one
1043: of these is the correct bounce solution.The bounce solution is the 
1044: real solution to the classical Euclidean equation of motion. The M2 
1045: solution satisfies the bounce criteria and is given by 
1046: (in the Euclidean form)
1047: \ba
1048: \label{bouncesol}
1049: a(\tau) = \frac{1}{\sqrt{2\lambda}} \sqrt{\left( 1 
1050: + \sqrt{\left( 1 - \frac{4 \nu}{\lambda}\right)} \cos 
1051: (2 \sqrt{\lambda}\tau) \right)}
1052: \ea
1053: This solution is the modified bounce and has some interesting 
1054: features.
1055: 
1056: \begin{figure}
1057: \begin{center}
1058: \epsfig{file=barrel.eps, width=3cm}
1059: \vspace{0.1in}
1060: \caption{The Modified Bounce : It is a squashed version of $S^4$ resulting in a ``barrel''.}
1061: \label{fig6}
1062: \end{center}
1063: \end{figure}
1064: 
1065: \begin{itemize}
1066: \item
1067: It is a deformation of the $S^4$ instanton from spherical to barrel-shaped,
1068: with $ - \pi/2 \sqrt{\lambda} \le \tau \le  \pi/2 \sqrt{\lambda}$.
1069: It reduces to the usual $S^4$ instanton when $\nu = 0$, as expected
1070: \ba
1071: a(\tau) \to \frac{1}{\sqrt{\lambda}} \cos (\sqrt{\lambda} \tau)
1072: \ea
1073: Here $\nu$ is the deformation parameter.
1074: \item
1075: For large values of $\nu$ the bounce is destroyed. Presumably, there 
1076: is no more
1077: quantum tunneling because of excessive decoherence. This critical
1078: value of $\lambda$ is given by $\lambda_c= 4 \nu $. 
1079: For tunneling $\lambda > \lambda_c$.
1080: 
1081: %That the deformed $S^4$ has to do with the lifting of ``nothing''  
1082: %can also be seen by noting that the modified bounce corresponds
1083: %to a modified Wheeler-DeWitt equation or, what is equivalent, a
1084: %modified Hamiltonian constraint. This, and the consequence this has
1085: %for the boundedness of the gravitational potential, we explain in the next section. 
1086: 
1087: \end{itemize}
1088: 
1089: \begin{figure}
1090: \begin{center}
1091: \epsfig{file=graphs.eps, width=9cm}
1092: \vspace{0.1in}
1093: \caption{The solid curve is $F= - S_{E,dC}$, 
1094: using the exact value of $\Im$. The dotted curve shows $F$ using 
1095: $\Im \simeq 8/3$. The light flat curve is the 
1096: integral $\Im$, which asymptotes to $8/3$ for large $\lambda$. 
1097: The plots are given with $M_{Pl}/M_s = 10^3$.
1098: The dotted and the solid curves differ only slightly, 
1099: justifying the approximation made.}
1100: %{The solid curve is the negative modified action, $-F$, 
1101: %(Eq.(\ref{modac})) using the exact value of $\Im$ (Eq.(\ref{im1})). 
1102: %The dotted curve shows $-F$ using approximation in Eq.(\ref{im2}). 
1103: %The light flat curve shows the shape of the integral $\Im$. This 
1104: %justifies the approximation made.}
1105: \label{fig7}
1106: \end{center}
1107: \end{figure}
1108: 
1109: The modified Euclidean action can now be calculated. One has to be
1110: careful though to include the O$({\nu}/{\lambda^2})$ contribution
1111: from the $-a\dot{a}^2 - a + \lambda a^3$ part. This contribution
1112:  will add to the O$({\nu}/{\lambda^2})$ contribution from the
1113: $\frac{\nu}{\lambda^2 a(\tau)}$ term and give the total O$({\nu}/{\lambda^2})$ 
1114: modification. We state the final result
1115: \ba
1116: \label{modac}
1117: -F =S_{E,dC} =\left( -\frac{1}{4\lambda} + 
1118: \frac{\nu}{\lambda^2}\right) \Im
1119: \ea 
1120: where $\Im$ is an integral given by
1121: \ba
1122: \label{im1}
1123: \Im = \int_{\frac{-\pi}{2\sqrt{\lambda}}}^{\frac{\pi}
1124: {2\sqrt{\lambda}}}d\tau \left( \frac{\sin^2(2\sqrt
1125: {\lambda}\tau)}{a(\tau)} \right)
1126: \ea
1127:  where $a(\tau)$ is the bounce solution given by Eq.(\ref{bouncesol}). 
1128:  The exact value of $\Im$ involves
1129: an elliptic integral, but we can make a fairly accurate 
1130: estimate and the result is
1131: \ba
1132: \label{im2}
1133: \Im \simeq \frac{8}{3}
1134: \ea
1135: 
1136: For tunneling from nothing, we should consider the barrel to be a deformed $S^4$ with 
1137: the same topology. That is, we should include the contribution of end plates of the barrel 
1138: in the evaluation of $S_{E,dC}$ in Eq.(\ref{modac}). At the end plates, 
1139: $$ a = \frac{1}{\sqrt{2 \lambda}} \left(1-\sqrt{1 - \frac{4\nu}{\lambda}}\right)^{1/2} > 0$$
1140: However, their contribution to $S_{E,dC}$ is zero because the weighing factor $\sin^2(2\sqrt
1141: {\lambda}\tau)$ in Eq.(\ref{im1}) vanishes at the end plates.
1142: 
1143: As expected, for $\nu = 0$ we get the usual Hartle-Hawking
1144: wavefunction. Now, including the environment,
1145: %Up to a O$({\nu}/{\lambda^2})$, however, the 
1146: the tunneling probability is given by 
1147: %modified wavefunction is just given by
1148: \ba
1149: \label{modwf}
1150: P \simeq e ^{\left(\frac{2}{3\lambda} - \frac{8\nu}{3\lambda^2}\right)}
1151: \ea
1152: For $\lambda \to 0$, higher order decoherence effects should be included.
1153: These effects can come from a more careful treatment of the leading term
1154: we have obtained, or from higher order interaction of the modes with $a$
1155: and among themselves. Quantum effects may become relevant. This is 
1156: especially so at large $\lambda$. Note that $S_{E,dC}$ peaks around
1157: $\lambda = 8 \nu$. So for regions with non-zero tunneling probability, 
1158: $\lambda > \lambda_{c} = 4\nu$. 
1159: 
1160: 
1161: 
1162: \section{The Modified Hamilton Constraint and the Wheeler-DeWitt Equation}
1163: 
1164: The modified action, as given in Eq.(\ref{bounce}), leads to a modified
1165: Hamiltonian constraint. The modified Lorentzian action is given by
1166: \ba
1167: S = \frac{1}{2}\int d\tau \left( - a\dot{a}^2 + a - \lambda a^3 
1168: - \frac{\nu}{\lambda^2 a} \right)
1169: \ea 
1170: The modified Hamiltonian constraint is
1171: \ba
1172: H = \frac{1}{2a}\left( -\Pi_{a}^2 - a^2 + \lambda a^4  
1173: + \frac{\nu}{\lambda^2} \right) = 0
1174: \ea
1175: where $\Pi_a = -a\dot{a}$ is the conjugate momentum. 
1176: Using the Hamiltonian constraint and the equation of motion following from the modified 
1177: Lorentzian action, we have 
1178: \ba
1179: (\frac{\dot a}{a})^2 + \frac{1}{a^2} = \lambda + \frac{\nu}{\lambda^2 a^4} \\ \nonumber
1180: \frac{\ddot a}{a} = \lambda - \frac{\nu}{\lambda^2 a^4}
1181: \ea
1182: Since ${\ddot a}/a= - \sum (\rho_i +3p_i)/2=-\sum \rho_i (1 +3\omega_i)/2$, this implies that 
1183: the new $\nu$ term has equation of state $\omega=1/3$, precisely that of ordinary radiation. 
1184: 
1185: One can write down a differential equation describing the evolution
1186: of the wavefunction - the Wheeler-DeWitt (WdW) equation -by imposing 
1187: the condition as an operator equation. This is quantum gravitational analog of
1188: the Schrodinger equation in quantum mechanics.
1189: \ba
1190: \label{wdw}
1191: -\hat{H}\Psi =
1192: % 0 \nonumber \\
1193: \left( \frac{\hat{\Pi_{a}}^2}{2} + U(a)\right)\Psi = 0 \nonumber 
1194: \ea
1195: where $\hat{\Pi_a} = -i\partial / \partial a$ and $U(a)$ is 
1196: the gravitational potential given by
1197: \ba
1198: U(a) = U_0(a)  - \frac{\nu}{\lambda^2} = a^{2} - \lambda a^4  - \frac{\nu}{\lambda^2}
1199: \ea
1200: Recall the case without the $\nu$ term. 
1201: In the classically forbidden region, i.e., the under-barrier region ${\lambda}^{-1/2} \ge a \ge 0$, 
1202: the WKB solutions for the tunneling amplitude from $a=0$ to $a=\lambda^{-1/2}$ are
1203: \ba
1204: \label{HKWKB}
1205:  \Psi_{\pm} \simeq e^{\pm \int_0^{1/\sqrt{\lambda}} |\hat \Pi(a')| da'}
1206: \ea
1207: %Because the kinetic term in Hamiltonian $\hat H$ is negative. 
1208: Note the Hartle-Hawking no boundary 
1209: prescription requires us to take the positive sign in the exponent in  Eq.(\ref{HKWKB}),
1210: that is, $\Psi_{+}$. This yields the HH wavefunction. 
1211: Including the $\nu$ term is like solving the wave equation 
1212: with a positive (instead of zero) energy eigenvalue $E=\nu/\lambda^2$
1213: (see Figure (\ref{fig8})), 
1214: \ba
1215: \left( \frac{\hat{\Pi_{a}}^2}{2} + U_{0}(a)\right)\Psi(a) = E \Psi (a)   \nonumber 
1216: \ea
1217: The presence of $E>0$ decreases the tunneling amplitude $\Psi_{+}$.
1218: \begin{figure}
1219: \begin{center}
1220: \epsfig{file=pot.eps, width=9cm}
1221: \vspace{0.1in}
1222: \caption{The solid curve is $U_{0}(a)=a^2 -\lambda a^{4}$.
1223: The effect of the new term is simply to raise the energy eigenvalue from $0$ to $E=\nu/\lambda^2>0$.} 
1224: \label{fig8}
1225: \end{center}
1226: \end{figure}
1227: To see the origin of $E$, consider the presence of a generic field $\chi$, so the WdW equation is
1228: crudely given by
1229: \ba
1230: \label{wd}
1231: \left(-\frac{\hat{\Pi_{a}}^2}{2} - U_{0}(a) + \frac{\Pi_{\chi}^2}{2} + \omega (\chi, a)^{2} \right) \Psi(a, \chi) = 0
1232: \ea
1233: Note that the kinetic terms in Eq.(\ref{wd}) have opposite signs. Let 
1234: \ba
1235: \Psi(a, \chi) = \Sigma_{n} \psi_{n}(a) u_{n}(\chi)
1236: \ea
1237: then the above equation can be separated,
1238: \ba
1239: \label{twoeq}
1240: \left(\frac{\Pi_{\chi}^2}{2} + \omega (\chi, a)^{2} \right) u_{n}(\chi)
1241: = \epsilon_{n}(a)u_{n}(\chi) \\ \nonumber
1242: \left(-\frac{1}{2}\frac{d^2\psi_{n}}{da^2} + U_{0}(a)\psi_{n}\right)
1243: = \epsilon_{n}(a)  \psi_{n} 
1244: \ea
1245: Taking the ground state, we obtain the zero point energy  $\epsilon_{0}$, with $\Psi(a)=\psi_{0}(a)$. Including the zero point energy of all $\chi$ fields then yields $E$, which turns out to be
1246:  independent of $a$. 
1247: 
1248: If instead, we take $\Psi_{-}$ in Eq.(\ref{HKWKB}), as suggested by 
1249: Linde and Vilenkin \cite{Linde:1984mx} and which is more familiar in quantum mechanics, 
1250: we see that the inclusion of the $\nu$ term enhances tunneling. 
1251: This means interaction with the environment enhances the tunneling amplitude, which is 
1252: counter-intuitive. Furthermore, the highest tunneling probability will go for large $\Lambda$, 
1253: when the semi-classical approximation used here breaks down. 
1254: %To be consistent with ordinary quantum field theory, 
1255: We believe the Hartle-Hawking no boundary prescription is the correct one for the
1256: spontaneous creation of the universe.
1257: 
1258: As we see clearly now, this $\nu$ term plays the role of radiation.
1259: The deformed $S^4$ has, in a sense, allow the spontaneous creation of a universe
1260: with some radiation in it. In pure gravity, this is simply gravitational radiation.
1261: Due to the presence of such radiation, the cosmic
1262: scale factor $a(\tau)$ does not vanish any more at the poles of $S^4$.
1263: In fact, the decoherence has led to a flattening of the two poles.
1264: 
1265: We note that,
1266: in theories with dynamical $\lambda$, there will be momentum terms
1267: associated with $\lambda$ in the Wheeler-DeWitt equation. For example,
1268: if $\lambda$ is associated with the potential energy of some scalar
1269: field $\phi$, then there will be momentum term $\hat{\Pi_{\phi}}$ in
1270: the Wheeler-DeWitt equation. However, the inclusion of such a momentum
1271: term for $\lambda$ is not expected to change the boundedness of the
1272: gravitational potential.
1273:  
1274: 
1275: For large cosmological constant $\Lambda$, the radiation is suppressed. In this case, one 
1276: may ignore it. However, as $\Lambda$ decreases and the size of the universe grows, radiation 
1277: becomes important. More radiation also means stronger suppression of the tunneling, consistent
1278: with the intuitive picture of tunneling while interacting with the environment. It is this suppression that provides a bound to the Euclidean gravity action.
1279:  
1280: 
1281: %To evaluate the above expression further we need to know the number of modes 
1282: %that contribute to the decoherence. There are two natural cut-offs in the 
1283: %problem. The infrared cut-off comes from the Hubble
1284: %length $H^{-1}$. We do not expect larger wavelenths to play any role 
1285: %in the problem. The ultra-violet cut-off comes from the string length 
1286: %$l_{s}$. Since the substcript ``$n$''  really is a shorthand notation
1287: %for ${n,l,m}$, the number of modes between the cutoff scales 
1288: %$H^{-1}$ and $l_{s}$ are :
1289: 
1290: %\baray
1291: %N = \left ( \frac{H^{-1}}{l_{s}} \right ) ^{3}
1292: %\earay
1293: 
1294: %As $N$ will be a large number and $\omega_{n}(\alpha)$ will 
1295: %quickly become much larger than $T^{-1}$, therefore, 
1296: %we can approximate Eq.(\ref{dec}) as
1297: 
1298: %\baray
1299: %\Psi[\alpha] = \int D[\alpha]  
1300: %e^{-I^{E}_{o}[\alpha]} \prod_{n \geq 1} e^{-\omega_{n}(\alpha)T/2}
1301: %\earay
1302: %Doing a steepest descent calculation, this leads to
1303: %\baray
1304: %\Psi[\alpha] = e^{\frac{3\pi}{G\Lambda} - NT <\omega>/2}
1305: %\earay
1306: %where $<\omega>$ is the average frequency given by
1307: %\baray
1308: %<\omega> = \frac{1}{N} \int_{0}^{N} dn \omega(n,\alpha) = \frac{H N}{2}
1309: %\earay
1310: 
1311: \section{Connection to SOUP}
1312: 
1313: In \cite{Firouzjahi:2004mx} we lay out the motivation for SOUP, short
1314: for ``Selection of the Original Universe Principle''. This is an alternate
1315: to the Anthropic Principle.  Since observational evidence of an 
1316: inflationary epoch is very strong,
1317: we suggest that the selection of our particular vacuum state 
1318: follows from the evolution of the inflationary epoch.
1319: That is, our particular vacuum site in the cosmic landscape
1320: must be at the end of a road that an 
1321: inflationary universe will naturally follow. Any vacuum state that
1322: cannot be reached by (or connected to) an inflationary stage can be 
1323: ignored in the search of candidate vacua. That is, the issue of the
1324: selection of our vacuum state becomes the question of the selection 
1325: of an inflationary universe, or the selection of an original universe 
1326: that eventually evolves to an inflationary universe, which then 
1327: evolves to our universe today.
1328: The landscape of inflationary states/universes should be much better 
1329: under control, since the inflationary scale is rather close to
1330: the string scale.
1331: In \cite{Firouzjahi:2004mx} we proposed that, by analyzing all 
1332: known string vacua and string inflationary scenarios, one may 
1333: be able to pin down SOUP. Here, the rate 
1334: of tunneling from nothing (i.e., no classical spacetime) to a 
1335: deSitter universe (much like the inflationary universe) is now
1336: given by $\Gamma \simeq \exp(F)= \exp(-S_{E,dC})$, where
1337: \ba
1338: \label{FS1}
1339: - F = S_{E,dC} = S_{E,0}+ D \simeq -\frac{3 \pi}{G\Lambda} + 
1340: \frac{n_{dof}}{4 \pi^2} \frac{V_4}{l_s^4}
1341: \ea
1342: where $D$ is the decoherence term, $l_s$ is the cut-off scale and $n_{dof}$ is the number of light degrees of freedom.  Here $V_4$ is the 4-volume of the 
1343: instanton. For $S^4$, $V_4 = 8 \pi^2 r^4/3= 24 \pi^2/{\Lambda^2}$ is its area.
1344: In string theory, that cut-off is naturally provided by the string scale.
1345: Note that $S_{E,dC}$ is now bounded from below. 
1346: 
1347: To gain some idea of the magnitude of $F$, we use Eq.(\ref{FS}) to crudely estimate the value of $\Lambda$ with maximum tunneling probability,
1348: \ba
1349: \Lambda_{max} \simeq \frac{4 n_{dof} G}{\pi l_s^4} \nonumber
1350: \ea
1351: so that
1352: \ba
1353:  F_{max} \simeq \frac{3}{2n_{dof}} (2 \pi M_{Pl}  l_s)^4 \nonumber
1354: \ea
1355: For the string scale a few orders of magnitude smaller than $M_{Pl}$, we see that $\Lambda_{max}$ takes a value quite close to that expected in an inflationary universe,
1356: with $F \sim 10^{12}$. On the other hand, the critical value of $\Lambda$ is $\Lambda_c \simeq \Lambda_{max}/2$. For $\Lambda < \Lambda_c$, the barrel-shaped 
1357: instanton is destroyed. At $\Lambda_c$,
1358: $F(\Lambda_c)\simeq 0$, so the tunneling probability at $\Lambda$ close to $\Lambda_c$ is already negligibly small compared to that at $\Lambda_{max}$.
1359: Tunneling to a supersymmetric vacuum is totally suppressed.
1360: For $l_s$ around the $TeV$ scale \cite{Arkani-Hamed:1998rs}, we find that 
1361: $\Lambda_{max}$ is quite close to today's dark energy value. This is similar in spirit to Ref\cite{Hsu:2004jt}. 
1362: However, this scenario will imply that our universe has not really gone through the 
1363: whole hot big bang and certainly not the standard inflationary epoch.
1364: 
1365: The key tool is the modified bounce. Although we only deal with the 
1366: $S^4$ instanton and its modification 
1367: in this paper, one can easily generalize this analysis to higher dimensions.
1368: We expect the form of the tunneling probability to be 
1369: $P \sim e^F$, with $F = -S_{E,dC}= -S_E -D$. In $10$-D, 
1370: \ba
1371: \label{new10S}
1372: S_{E,dC} \simeq S_{E,10}  + c \left(\frac{V_{10}}{l_s^{10}}\right)
1373: \ea
1374: where $S_{E,10}$ is the 10-D Euclidean action determined in mini-superspace and 
1375: $V_{10}$ is the 10-dimensional volume of the instanton. 
1376: In effective $4$-D theory, $S_{E,10}$ reduces to $-{3 \pi}/{G\Lambda}$,
1377: since 
1378: $$8 \pi G = M_{Pl}^{-2}=\frac{g_s^2 l_s^8}{4 \pi V_6}$$ 
1379: where $V_6$ is the 6-D compactification volume.
1380: (Recall $\alpha^{\prime}=M_s^{-2}=(l_s/2 \pi)^2$.)
1381: The constant $c$ has to be calculated for each vacuum and will
1382: depend on the details of the vacuum.
1383: To get an order of magnitude estimate of $c$ we have, by comparing to the 4-D case,
1384: \ba
1385: c \simeq \frac{n_{dof}}{ \pi} \frac{1}{M_{Pl}^2 l_s^2 g_s^2}
1386: \ea
1387: so we expect $c$ to be small.
1388: In \cite{Firouzjahi:2004mx} we have considered such instantons. We did
1389: not do a careful calculation of $\nu$ then, that has been the main 
1390: content of the present paper, but we phenomenologically guesstimated a 
1391: possible range for its values and calculated the creation probabilities for various
1392: vacua in the cosmic landscape. 
1393:  Among all the known string vacua, we find that the universe 
1394: most likely to be created is a KKLMMT type inflationary vacuum. 
1395: (For details, we refer the reader
1396: to \cite{Firouzjahi:2004mx}.) The probabilities calculated were very
1397: robust and did not depend on the exact value of $c$. 
1398: Changing from $V_9$ to $V_{10}$ does not change the overall qualitative results. 
1399: Therefore, we expect the conclusions to stay the same.
1400: 
1401: There are some minor differences.
1402: There we find that tunneling to a universe with today's dark energy is very much suppressed.
1403: Here we see that the cosmological constant corresponding to today's dark energy is actually 
1404: below the critical value, so tunneling directly to today's universe is simply zero.
1405: A more careful calculation based on Eq.(\ref{new10S}) is clearly needed.
1406: However, the important point
1407: is that one can calculate it, at least in principle, for any
1408: given vacuum and, therefore, compare their probabilities. This would
1409: be a huge improvement over the Anthropic Principle. \\
1410: 
1411:  One can calculate the tunnelling probability to $S^{10}$ using 
1412: Eq.(\ref{new10S}). For $S^{10}$, $S_{E,10} = \frac{\Lambda V_{10}}
1413: {32 \pi G_{10}}$, where $V_{10} = \frac{2^{11} 5! \pi^5}{10!}/
1414: {H^{10}}$, is the volume of $S^{10}$. Note that for $S^{10}$,
1415: $H^2 = \Lambda/ 36$. Without branes, the solution we consider is a 10D supersymmetric vacuum.
1416: Tunneling to this vacuum is suppressed. Next we consider tunneling to a deSitter-like vacuum. 
1417: Let there be $N$ $D9$ brane-antibrane pairs
1418: in the $S^{10}$. So 
1419: \ba
1420: \Lambda = 2N \times 8\pi G_{10} \times \frac{M_s^{10}}{(2\pi)^9 g_s}
1421: = \frac{g_s M_s^2 N}{(2\pi)^4}
1422: \ea
1423: Maximizing $F$ in terms of $N$ gives
1424: \ba
1425: N = \frac{5 n_{dof}(2\pi)^6}{g_s}\left( \frac{M_s}{M_{Pl}}\right)^2
1426: \ea
1427: This means $N = 1$ is a reasonable choice for
1428: %$n_{dof} = 2$, 
1429: $g_s \sim 1$, and $M_s/M_{Pl} = 10^{-3}$. One can then calculate $F$ for $N = 1$ for $S^{10}$. 
1430: In this case, $F$ is dominated by $S_{E,0}$. One gets 
1431: \ba
1432: F(S^{10}) \simeq -S_{E,0}(S^{10}) = \frac{2^{11}~ 5!~ (36 \pi)^5}{10!} \simeq 4 \times 10^{9}
1433: \ea
1434: 
1435: Following Ref\cite{Firouzjahi:2004mx}, we see that other related geometries such as $S^5 \times S^5$, $S^4 \times S^6$, $S^4 \times S^3 \times S^3$ etc. have slightly smaller but similar values of $F$.
1436: 
1437: For a KKLMMT-like inflationary universe, the choice of fluxes fixes the string scale as well as 
1438: the scale of inflation, where $G$ and the density perturbation measured in the cosmic microwave background radiation are used as input parameters. Using Eq.(\ref{FS1}) and $\Lambda = 8\pi G \rho_{vac}$, where the compactified volume has been absorbed into $G$, 
1439: one can rewrite Eq.(\ref{FS1}) so
1440: \ba
1441: F = \frac{3}{8}\left( \frac{\sqrt{8\pi}M_{Pl}}{\rho_{vac}^{1/4}}\right)^4 - \frac{3}{4 (2\pi)^6} \left( \frac{M_{s}}{\sqrt{8\pi}M_{Pl}}\right)^4 \left( \frac{\sqrt{8\pi}M_{Pl}}{\rho_{vac}^{1/4}}\right)^8
1442: \ea
1443: Plugging in the above mentioned values for the mass scales, we get $F$. 
1444: For a typical choice of fluxes that would give $M_s \sim 10^{15}$GeV and an inflationary scale $\rho_{vac}^{1/4} \sim 10^{14}$GeV, with the Planck mass $M_{Pl} \sim 10^{18}$GeV, one gets
1445:  $$F \sim 10^{16}$$
1446:  for a realistic brane inflationary scenario.
1447: This big increase in $F$ from the 10D scenario is a result of the decrease in the effective cosmological constant due to the warping of the geometry. By varying the RR and the NS-NS fluxes, one can maximize $F$ in a way which is not possible in 10D.  Of course, this value is sensitive to the details of the model. A more careful calculation will be important.
1448: 
1449: \section{Discussion}
1450: 
1451: Let us make some comments here.
1452: 
1453: \itemize
1454: 
1455: \item The loop correction to the Euclidean action has been calculated before
1456: \cite{Hawking:1976ja,Gibbons:1978ji}. It has the form
1457: $\log(\Lambda/\mu)$, so $S_E$ becomes 
1458: $$S_E= - \frac{3 \pi}{G\Lambda}  \to  - \frac{3 \pi}{G\Lambda}\left(1 + \beta G\Lambda\log(\Lambda/\mu)\right)$$
1459: where $\beta$ measures the number of fields involved.
1460: Ref\cite{Barvinsky:1992dz} considers the metric perturbation also; however, 
1461: they then relate it to the one-loop contribution, yielding the above result.
1462: Note that this loop correction does not solve the boundedness problem in the
1463: HH wavefunction. Its implications on probability of inflation has also been
1464: discussed (see \cite{Barvinsky:1996ce} and the references therein). 
1465:  
1466: \item Decoherence effects are to be distinguished from the 
1467: particle creation effects that have been discussed in the literature
1468: \cite{Rubakov:1984bh}.
1469: Particle creation effects arise at the one-loop level, while the
1470: decoherence effect discussed here is a back-reaction improved quantum correction. 
1471: 
1472: \item Back reaction is crucial in obtaining the decoherence term. This reminds one of the 
1473: situation in quantum field theory, where
1474: a simple 1-loop correction to the coupling takes on new significance when it is 
1475: renormalization group improved. 
1476: 
1477: \item Integrating out the environment provides a new interaction term for
1478: the cosmic scale factor $a$ in the Einstein equation. This decoherence term 
1479: behaves just like ordinary radiation and generates an effective term in the Euclidean action for the instanton. 
1480: The presence of this term provides the lower bound to the resulting Euclidean gravity.
1481: 
1482: %\item It is interesting to note that in terms of the string coupling $g_s$, an
1483: %O$({\nu}/{\lambda^2})$ correction to the Euclidean
1484: %action corresponds to an order O$({g_s^{2}})$ correction.
1485: %This is a novel feature. ?
1486: 
1487: \item The correction to the Hartle-Hawking wavefunction 
1488: results in a decrease in the Euclidean action. If we are
1489: to interpret the Euclidean action as the entropy, then
1490: there is a decrease in the entropy due to decoherence.
1491: This is in accordance with the Bekenstein bound on entropy.
1492: A pure deSitter space provides the upper bound on entropy.
1493: Decoherence, which is due to the inclusion of extra degrees
1494: of freedom, then leads to a lower entropy respecting the entropy bound.
1495: 
1496: \item Once the inflationary universe is created, inflation simply red-shifts 
1497: the radiation so the radiation term quickly becomes negligible.  
1498: It seems that the initial radiation is present just to suppress the tunneling.
1499: 
1500: \item It is interesting to speculate what happens after inflation.
1501: Towards the end of inflation, the inflaton field (or the associated
1502: tachyon mode) rolls down to the bottom of the inflaton potential. The
1503: universe is expected to be heated up to start the hot big bang epoch.
1504: It is not unreasonable to expect the wavefunction of the universe to be
1505: a linear superposition of many vacua (say, a billion of them) at the
1506: foothill of the inflaton potential. This is like a Bloch wavefunction
1507: as proposed in Ref\cite{Kane:2004ct}. Since tunneling happens only
1508: between vacua with (semi-)positive cosmological constants, this
1509: wavefunction should be a superposition of vacua with only positive
1510: cosmological constants. The ground state energy of this wavefunction is
1511: expected to be much smaller than the average vacuum energy, a property
1512: of Bloch wavefunctions. As the universe evolves, tunneling among the
1513: vacua will become suppressed, due to cooling as well as decoherence.
1514: Eventually, the environment collapses the wavefunction to a single
1515: vacuum. Since the Bloch wavefunction is able to sample many vacua, it
1516: is natural to expect that it will collapse to the vacuum state with the
1517: smallest positive vacuum energy within teh sample. This may partly
1518: explain why the dark energy is so small.
1519: 
1520: \item In terms of recent work on the statistics of the landscape 
1521: \cite{Kachru:2003aw,Douglas:2003um},
1522: one may view our work as providing a measure to the counting
1523: of vacua. Each vacuum is weighted with its probability of tunneling 
1524: from nothing. Since the exponent of the tunneling probability can 
1525: differ by many orders of magnitude, this is likely to be the dominant 
1526: contribution to the measure. We argue that a 4-dimensional inflationary 
1527: universe very much like ours has large probability; this implies that vacua
1528: that cannot be reached after inflation should have vanishing measure.
1529: 
1530: \item The effect of decoherence on other tunneling problems in quantum gravity
1531: should be re-examined. These include tunneling in eternal inflation, the 
1532: Coleman-deLuccia tunneling etc.
1533: 
1534: 
1535: \section{Summary and Remarks}
1536:      
1537: We address a number of questions in this paper.
1538: 
1539: First we ask what happens to the well known
1540: $S^4$ bounce solution to the Euclidean Einstein equations 
1541: when the perturbations to the metric are taken into account. 
1542: How does the bounce change as a result?
1543: In particular, is there a parameter
1544: characterizing the perturbations that describes the modification
1545: of the bounce? Is there a range of this parameter that destroys 
1546: the bounce altogether? 
1547: 
1548: To answer the first question, we apply the path integral
1549: techniques to trace out the perturbations. The result is summarized in the introduction.
1550: Taking back-reaction into account, the Euclidean action now includes an additional term that encapsulates
1551: the effect of the perturbations. Depending on the value of the 
1552: parameter $c$ or equivalently $\nu$, the bounce gets deformed. For a fairly
1553: large range of this parameter, the effect is to just deform the bounce
1554: solution. Since this contribution is always positive, the tunneling
1555: is always suppressed. At a critical value, the bounce is destroyed. One 
1556: may interpret that tunneling is forbidden in this limit. 
1557: 
1558: Next, we ask what these perturbations do to the
1559: boundedness of the Euclidean gravitational action. As is well known,
1560: the Euclidean gravitational action for a closed spacetime is not
1561: bounded from below for theories with a dynamical cosmological constant. 
1562: This manifests, for example, as the infinite peaking of
1563: the wavefunction of the universe at the vanishing value of $\Lambda$.  
1564: Does the inclusion of the perturbations change anything here? 
1565: 
1566: To answer this question, we see that the effect of the perturbations is
1567: to change the wavefunction of the universe from $\exp (3\pi/G\Lambda)$
1568: to $\exp (3\pi/G\Lambda - C/\Lambda^2)$ where $c$ depends on the 
1569: particular perturbations that we are looking at. So, at least for
1570: the problem at hand, the inclusion of the perturbations and their back reaction makes the
1571: Euclidean action bounded from below. This renders the wavefunction 
1572: normalizable. In the study of tunneling, it makes no sense
1573: to talk about gravity without taking into account the perturbations 
1574: to the metric. In this sense, quantum gravity is consistent as long as 
1575: we are careful to include the effects of the perturbative modes which are 
1576: always present.
1577: 
1578: Once we have a sensible wavefunction, we can now go ahead and 
1579: apply it to the cosmic landscape in string theory. Since Euclidean action and 
1580: tunneling probability are both dimensionless, we can compare the tunneling 
1581: from nothing to any point in the landscape and find the sites that have the largest 
1582: probability. This program hopefully will allow us to understand why we are 
1583: where we are, without resorting to the anthropic principle.
1584: 
1585:   
1586: \vspace{0.5cm}
1587: 
1588: {\large{\bf{Acknowledgments}}}\\
1589: 
1590: We thank Faisal Ahmad, Andrei Barvinski, Spencer Chang, Jacques Distler, 
1591: Hassan Firouzjahi, Lerrain Friedel, Jim Hartle, Gordy Kane, Louis Leblond, Andrei Linde, 
1592: Juan Maldacena, Gautam Mandal, Liam McAllister, Lubos Motl, 
1593: Hirosi Ooguri, Koenraad Schalm,  Jan Pieter van der Schaar, Jim Sethna,  Sarah Shandera, Gary Shiu, Ben Shlaer
1594: and  Cumrun Vafa for useful discussions. This work is supported by the National Science Foundation under Grant No. PHY-009831.
1595: 
1596: \vspace{0.5cm}
1597: 
1598: \appendix
1599: 
1600: \section{A Quantum Mechanical Example}
1601: 
1602: To see the basic idea, recall the quantum tunneling of the system
1603: \ba
1604: L_0= \frac{M}{2} {\dot q}^2 -V(q)
1605: \ea
1606: with a quartic $V(q)$ as shown in Fig. 2. The tunneling rate 
1607: $\Gamma$ from the local minimum at $q=0$ to the exit point 
1608: $q=q_0$ is well-known,
1609: \ba
1610: \Gamma &=& A \exp (-S_0) \nonumber \\
1611: S_0 &=& \int^{q_0} \sqrt{2MV(q)} dq =\int  \left 
1612: (\frac{M}{2} {\dot q}^2 +V(q) \right) d\tau
1613: \ea
1614: where $\tau$ is the Euclidean time and $S_0$ is the bounce, 
1615: i.e., the instanton solution \cite{Coleman:1977py}.
1616: This WKB approximation is good provided that the height of 
1617: the barrier is larger than $\om_0$, where 
1618: $$M \om_0^2 =  \frac{\partial^2 V}{\partial q^2}|_{q=0}$$
1619: Note that, for $V(q)$ bounded from below, $S_0$ is bounded 
1620: from below, as required by consistency.
1621: In a more realistic situation, the particle interacts with 
1622: the environment. Typically, this introduces a frictional force, 
1623: so the corresponding classical equation is given by
1624: \ba
1625: M{\ddot q} + \eta {\dot q} + \frac{\partial V}{\partial q} = 0
1626: \ea
1627: The impact of such a frictional term on the particle is to 
1628: suppress the tunneling rate. Consider the following system
1629: \ba
1630: L= \frac{M}{2} {\dot q}^2 -V(q) + \frac{1}{2} \sum_{\al} m_{\al} 
1631: \left({\dot x_{\al}}^2 - \om_{\al}^2 x_{\al}^2\right)
1632: -q \sum_{\al} C_{\al}x_{\al} -\frac{1}{2}M (\delta \om)^2 q^2
1633: \ea
1634: where one may consider $q$ to be the system and the $x_{\al}$ to 
1635: be the environment. The last term is a counter term introduced to 
1636: correct the shift in frequency,
1637: \ba 
1638: M (\delta \om)^2 &=&  \sum_{\al} \frac{C^2_{\al}}{m_{\al} \om^2_{\al}} 
1639: \ea
1640: The interactions of $q$ with the $x_{\al}$ introduces the friction term
1641: \ba
1642: \eta = \frac{\pi}{2} \sum_{\al} \frac{C_{\al}^2}{m_{\al} \om_{\al}^2} 
1643: \delta(\om_{\al }-\om) 
1644: \ea
1645: for $\om$ smaller than some critical $\om_c$.
1646: 
1647: The tunneling rate of this sytem can be easily found \cite{Sethna:1981dr,Caldeira:1982uj} 
1648: that the bounce $S_0$ increases to
1649: \ba
1650: S \simeq S_0 \left( 1 + \frac{\eta}{2M \om_0}\right)
1651: \ea
1652: That is, the interaction with the environment, or the friction, 
1653: suppresses the tunneling rate. This qualitative feature remains 
1654: true when the environment and its interaction with the
1655: system is more complicated.
1656: 
1657: 
1658: 
1659: \section{Calculations on the Squashed $S^4$}
1660: 
1661: This appendix gives details for Section $3$. We show how
1662: we perform the path integral over the $t_n$ mode.
1663: The solution to Eq.(\ref{t_n}) satisfying the initial condition 
1664: $t_n (x = 1 - \delta) = t_n^i$, and the final condition 
1665: $t_n (x = -1 + \delta) = t_n^f$,  is given by
1666: \ba
1667: \label{tnx}
1668: t_n(x) = \frac{F_n(x)}{\sqrt{1 - x^2}} 
1669: = \frac{A P^{-n}_{1}(x) + B Q^{n}_{1}(x)}{\sqrt{1 - x^2}}
1670: \ea 
1671: where $A$ and $B$ are given by
1672: \ba
1673: \label{AB}
1674: A = \sqrt{2\delta - \delta^2}\frac{t_n^i Q^n_{1}(-1 + \delta) - t_n^f 
1675: Q^n_{1}(1 - \delta)}{ P^{-n}_1(1 - \delta) Q^n_{1}(-1 + \delta) - 
1676:  P^{-n}_1(-1 + \delta) Q^n_{1}(1 - \delta)} \\ \nonumber
1677: B = - \sqrt{2\delta - \delta^2}\frac{t_n^i P^{-n}_{1}(-1 + \delta) - t_n^f 
1678: P^{-n}_{1}(1 - \delta)}{ P^{-n}_1(1 - \delta) Q^n_{1}(-1 + \delta) - 
1679:  P^{-n}_1(-1 + \delta) Q^n_{1}(1 - \delta)}
1680: \ea
1681: 
1682: The trace operation is defined as doing the following
1683: \ba
1684: \int dt_n^i \int dt_n^f \delta (t_n^i - t_n^f)\int^{(t_n^{initial} 
1685: = t_n^i)}_ {(t_n^{final} = t_n^f)}
1686: D[t_n] \exp \left( S^n_E [t_n(x)]\right)
1687: \ea
1688: 
1689: Since our purpose finally is to take to trace over the $t_n$ mode, we can
1690: set $t_n^i = t_n^f$ at this stage. The Euclidean action due to the $n$th 
1691: mode is then
1692: \ba
1693: S_{E}^{n} & = & \frac{1}{2} \left[a(\tau)^3 t_n \frac{dt_n}{d\tau}  
1694: \right]_{x = (1 - \delta)}^{x = (-1 + \delta)} \\ \nonumber
1695: & = & \frac{1}{2} \left[xF_n^2 + (1 -x^2)F_n\frac{dF_n}{dx}  
1696: \right]_{x = (1 - \delta)}^{x = (-1 + \delta)}
1697: \ea
1698: 
1699: Using Eq.(\ref{tnx}, \ref{AB}) and the symmetry properties of
1700: the associate Legendre functions
1701: \ba
1702: P^{-n}_1(-x) = - P^{-n}_1(x), &  \quad & Q^{n}_1(-x) = Q^{n}_1(x) \\ \nonumber
1703: \partial_{x}P^{-n}_1(-x) = \partial_{x}P^{-n}_1(-x), & \quad &
1704: \partial_{x}Q^{n}_1(-x) = - \partial_{x}Q^{n}_1(x)
1705: \ea
1706: one gets the following action
1707: \ba
1708: S_{E}^{n} = \delta(2 -\delta) \left[ (1 - \delta) + \delta(2 -\delta)
1709: \frac{\partial_{x}Q^{n}_1(1 -\delta)}{Q^{n}_1(1 -\delta)}\right] (t_n^i)^2
1710: \ea
1711: Next we must find the prefactor to the path integral for $t_n$.
1712: This can be found as explained in the Appendix C. The prefactor 
1713: is given by $\frac{1}{\sqrt{2\pi g_n(-1 + \delta)}}$, where $g_n(x)$
1714: is a solution to Eq.(\ref{t_n}), i.e.
1715: $ g_n(x)= \frac{C P^{-n}_{1}(x) + D Q^{n}_{1}(x)}{\sqrt{1 - x^2}}$, 
1716: and it satisfies the following conditions
1717: \ba
1718: g_n(1 -\delta) = 0, \quad \partial_x g_n(1 -\delta) = 1
1719: \ea
1720: It is easy to check that such a solution to Eq.(\ref{t_n}) 
1721: has the following values
1722: of $C$ and $D$
1723: \ba
1724: C = \sqrt{2\delta - \delta^2} \frac{Q^n_1(1-\delta)}{ Q^n_{1}(1 - \delta)
1725: \partial_x P^{-n}_1(1 - \delta) 
1726:  - P^{-n}_1(1 - \delta) \partial_x Q^n_{1}(1 - \delta)} \\ \nonumber
1727: D = - \sqrt{2\delta - \delta^2} \frac{P^{-n}_1(1-\delta)}{ Q^n_{1}(1 - \delta)
1728: \partial_x P^{-n}_1(1 - \delta) 
1729:  - P^{-n}_1(1 - \delta) \partial_x Q^n_{1}(1 - \delta)}
1730: \ea
1731: The prefactor is then given by
1732: \ba
1733: \frac{1}{\sqrt{2\pi g_n(-1 + \delta)}} = \frac{1}{\sqrt{4\pi}}
1734: \sqrt{\left( \frac{\partial_xQ^n_1(1-\delta)}{Q^n_1(1-\delta)} 
1735: - \frac{\partial_xP^{-n}_1(1-\delta)}{P^{-n}_1(1-\delta)} \right)}
1736: \\ \nonumber
1737: = \frac{1}{\sqrt{4\pi}} \frac{\sqrt{Q^n_1(1 -\delta) P^n_1(1 - \delta) (2\delta- \delta^2)}}{2^n} \sqrt{\frac{\Gamma\left( \frac{2-n}{2}\right) \Gamma\left( \frac{3-n}{2}\right)}{\Gamma\left( \frac{2+ n}{2}\right) \Gamma\left( \frac{3+n}{2}\right)} }
1738: \ea
1739: 
1740: 
1741: \section{Path Integral of an Oscillator with variable 
1742: mass and variable frequency}
1743: 
1744: Here we review some basic properties of path integral and apply them to the evaluation of the 
1745: prefactor in Eq.(\ref{PII}). To be concrete, we shall follow Ref\cite{Felsager:1981iy}.
1746: Consider an oscillator with mass $m(t)$ and frequency $\omega(t)$. What
1747: we have in mind, in particular, is a case like Eq.(\ref{PI}) which
1748: describes the path integral for a tensor mode (in Euclidean spacetime).
1749: Comparing Eq.(\ref{PI}) with this appendix, the time dependent mass
1750: would be $a(t)^3$ and the time dependent frequency would be
1751: ${(n^2 -1)}/{a^2}$ .  
1752: We would like to evaluate the propagator $K(x_f,T | x_i, 0)$. The action
1753: is given by
1754: \baray
1755: & S = \int_{o}^{T} dt \left( \frac{1}{2} m \left( \frac{dX}{dt}\right)^2 
1756: - \frac{1}{2} m \omega(t)^2 X^2 \right) 
1757: \earay
1758: The problem can be simplified by introducing a new "time" variable $u$ so as to map the 
1759: present problem to that of an oscillator with unit mass and variable frequency: 
1760: \baray
1761: \label{defnu}
1762: du = \frac{dt}{m(t)}
1763: \earay
1764: In terms of $u$ the action becomes
1765: \baray
1766: & S = \frac{1}{2} \int_{u_i}^{u_f} du \left( \left( \frac{dX}{du}\right)^2 
1767: - \Omega(u)^2 X^2 \right)
1768: \earay
1769: that is, the action for an oscillator with unit mass and variable 
1770: frequency $\Omega(u)$. As this is quadratic in $X(u)$, we can 
1771: expand around a classical solution, $X = x_{cl} + x$, where the classical 
1772: solution $x_{cl}(u)$ satisfies the equation of motion
1773: \ba
1774: \label{class_eqn}
1775: \frac{d^2 x_{cl}(u)}{du^2} + \Omega(u)^2 x_{cl}(u) = 0
1776: \ea
1777: where $\Omega(u)$ depends on $n$. For example,
1778: for the $n$th tensor field perturbation, $\Omega(u) = (n^2 -1)a(u)^4$. 
1779: One can use the above solution for $x_{cl}$, where $x_{cl}(u_i)=X(u_i)$ and 
1780: $x_{cl}(u_f)=X(u_f)$, to calculate the classical action $S[x_{cl}]$. This gives the
1781: saddle point value of the path integral. Following the discussion in Sec. 5, 
1782: the prefactor of the propagator is given by:
1783: \baray
1784: \label{prefactor}
1785: & \int_{x(u_i)=0}^{x(u_f)=0}D[x(t)] \exp \left[ \frac{i}{2} \int_{u_i}^{u_f}du
1786: \left(\left( \frac{dx}{du}\right)^2 - \Omega(u)^2 x^2  \right) \right] 
1787: \earay
1788: 
1789: By a further change of variables, we can change the action to a free
1790: particle action. To do so, let $g(u)$ be a solution of the 
1791: equation of motion (\ref{class_eqn})
1792: %\baray
1793: %\label{defng}
1794: %\frac{d^2 g(u)}{dt^2} + \Omega(u)^2 g(u) = 0
1795: %\earay
1796: such that $g(u)$ does not vanish at the initial end point
1797: \baray
1798: g(u_i) \neq 0
1799: \earay
1800: implying that $g(u)$ is not a path in Eq.(\ref{prefactor}).
1801: Define the following transformation of variables
1802: \baray
1803: \label{defny}
1804:  x(u) = g(u) \int_{u_i}^{u}\frac{y'(s)}{g(s)}ds=g(u)F(u)
1805: \earay
1806: where the function $y(u)$ obeys $y(u_i) = 0$. 
1807: Differentiating Eq.(\ref{defny}), one finds the inverse transformation
1808: \baray
1809: & y(u) = x(u) - \int_{u_i}^{u}\frac{g'(s)}{g(s)}x(s)ds
1810: \earay 
1811: In terms of the $y$ variable, we have:
1812: \baray
1813: & \left( \frac{d^2}{du^2} + \Omega(u)^2 \right)x(u) = \left(g''(u) 
1814: + \Omega(u)^2 g(u) \right) \int_{u_i}^{u}\left[ \frac{y'(s)}
1815: {g(s)}\right]ds \\ \nonumber
1816: & + \frac{g'(u)y'(u)}{g(u)} + y''(u) = \frac{g'(u)y'(u)}{g(u)} +   y''(u)
1817: \earay
1818: %But here the first term on the right vanishes due to Eq.(\ref{defng}). So,
1819: so the action becomes
1820: \baray
1821: & S[x(u)] = -\frac{1}{2} \int_{u_i}^{u_f}du \left[ F(u)g'(u)y'(u)+ 
1822: F(u)g(u)y''(u)\right]
1823: \earay
1824: %with $F(u) = \int_{u_i}^{u}ds\frac{y'(s)}{g(s)}$. 
1825: Keeping in mind that $x(u)$ vanishes at the boundaries (see Eq.(\ref{prefactor})), 
1826: a further partial integration then leads to
1827: \baray
1828: & S[x(t)] = \frac{1}{2}\int_{u_i}^{u_f}du \left[ \frac{dy}{du} \right]^2
1829: \earay
1830: This is just a free particle action, with boundary conditions
1831: %However the boundary condition for this free particle at $u = u_f$ is nonlocal
1832: \baray
1833: & y(u_i) = 0 ~~ ; \quad & y(u_f) = \int_{u_i}^{u_f}ds \frac{y'(s)}{g(s)} = 0
1834: \earay
1835: To impose the (non-local) boundary condition $x(u_f) = y(u_f)=0$, we introduce
1836: \baray
1837: & \delta(x(u_f)) = \frac{1}{2\pi} \int d\alpha ~ e^{-i \alpha x(u_f)}
1838: \earay
1839: The path integral can now be written as
1840: \baray
1841: &\int_{x(u_i)=0}^{x(u_f)=0} D[x(u)]\exp \left(iS[x(u)] \right)   
1842: = \frac{1}{2\pi} \int_{x(u_i)=0}^{x(u_f)~arbitrary} \int_{-\infty}^{\infty}
1843: d\alpha  D[y(u)] \nonumber \\  & \det \left[ \frac{\delta x}{\delta y}\right].
1844: \exp \left[-i \alpha g(u_f)\int_{u_i}^{u_f}ds 
1845: \frac{y'(s)}{g(s)} \right]. \exp \left[  \frac{i}{2}\int_{u_i}^{u_f}du
1846: \left( \frac{dy}{du}\right)^2\right]  
1847: \earay
1848: 
1849: As the transformation between $x(u)$ and $y(u)$ is linear, the Jacobian
1850: $\det \left[ \frac{\delta x}{\delta y}\right]$ is independent of $y(u)$.
1851: To carry out the above path integral, we just have to complete the 
1852: square by the use of the new variable
1853: \baray
1854: & \gamma(u) = y(u) - \alpha ug(u_f) \int_{u_i}^{u}\frac{ds}{g(s)}
1855: \earay
1856: The above path integral thus becomes
1857: \baray
1858:  ... &= \frac{1}{2\pi} \det \left[ \frac{\delta x}{\delta y}\right] 
1859: \int_{-\infty}^{\infty} d\alpha \exp \left[ -\frac{i}{2}\alpha^2 
1860: g(u_f)^2 \int_{u_i}^{u_f}\frac{du}{g(u)^2} \right] \\ \nonumber 
1861: & \int_{\gamma(u_i)=0}^
1862: {\gamma(u_f)arbitrary} D[\gamma(u)] \exp \left[ \frac{i}{2}
1863: \int_{u_i}^{u_f} \left( \frac{d\gamma}{du}\right)^2 \right]
1864: \earay
1865: 
1866: It is easy to carry out the $\alpha$ integral. To do the remaining 
1867: path integral, one just has to notice that it represents the 
1868: probability amplitude for finding the particle
1869: {\it anywhere} at the  time $u_f$. This probability is obviously $1$.
1870: \baray
1871: & \int_{\gamma(u_i)=0}^
1872: {\gamma(u_f)} D[\gamma(u)] \exp \left[ \frac{i}{2}
1873: \int_{u_i}^{u_f} \left( \frac{d\gamma}{du}\right)^2 \right]
1874: \\ \nonumber & = \int_{-\infty}^{\infty} dx K(x, u_f | 0, u_i) = 1 
1875: \earay
1876: where $\gamma(u_f)$ is arbitrary. Therefore, we have 
1877: \baray
1878: & \int_{x(u_i)=0}^{x(u_f)=0}D[x(t)] \exp \left[ \frac{i}{2} \int_{u_i}^{u_f}du
1879: \left(\left( \frac{dx}{du}\right)^2 - \Omega(u)^2 x^2  \right) \right] \\ \nonumber
1880: & = \det \left[ \frac{\delta x}{\delta y}\right] \sqrt{\frac{1}{2\pi 
1881: i g(u_f)^2 \int_{u_i}^{u_f}\frac{ds}{g(s)^2}}}
1882: \earay
1883: 
1884: It remains to find the value of the Jacobian. It can be found by a discretization
1885: process. The paths $x(u)$ and $g(u)$ can be replaced by the multidimensional points
1886: $(x_0, x_1, ..., x_N)$ and $(y_0, y_1,..., y_N)$ with $x_k = x(u_k)$ and
1887: $y_k = y(u_k)$. The linear transformation can then be approximated as
1888: \baray
1889: y_n & = & x_n - \frac{T}{N}\sum_{k=1}^{n}\frac{g'(u_k)}{g(u_k)} 
1890: \frac{x_k + x_{k-1}}{2} 
1891: \earay
1892: Then,
1893: \baray
1894:  J_N = \det \left[ \frac{\partial y_i}{\partial x_j} \right] 
1895:  = \prod_{k=1}^{N}\left( 1 - \frac{1}{2} \frac{g'(u_k)}{g(u_k) \frac{T}{N}} \right)
1896: \earay
1897: Taking the $N \to \infty$ limit, one gets
1898: \baray
1899:  \det \left[\frac{\delta y}{\delta x}\right] =  \exp \left[ -\frac{1}{2}
1900: \int_{u_i}^{u_f}du \frac{g'(u)}{g(u)} \right]  = 
1901: \sqrt{\frac{g(u_i)}{g(u_f)}} 
1902: \earay
1903: So the path integral prefactor becomes
1904: \baray
1905: \label{pref1}
1906: & \int_{x(u_i)=0}^{x(u_f)=0}D[x(t)] \exp \left[ \frac{i}{2} \int_{u_i}^{u_f}du
1907: \left(\left( \frac{dx}{du}\right)^2 - \Omega(u)^2 x^2  \right) \right] \\ \nonumber
1908: &=\Delta \left[\det \left( - \frac{d^2}{du^2} - \Omega(u)^2  \right) 
1909: \right]^{-1/2}  = \left({2\pi i g(u_i) g(u_f) \int_{u_i}^{u_f}\frac{ds}{g(s)^2}} \right)^{-1/2} 
1910: \earay
1911: where $\Delta$ is a normalization factor.
1912: One can check the validity of this result for simple cases like free particle
1913: and simple harmonic oscillator. 
1914: %We can simplify this result further. We can rewrite the path integral as:
1915: %\baray
1916: %& \int_{x(u_i)=0}^{x(u_f)=0}D[x(t)] \exp \left[ \frac{i}{2} \int_{u_i}^{u_f}du
1917: %x(u)\left(- \frac{d^2}{du^2} - \Omega(u)^2 \right)x(u) \right] \\ \nonumber
1918: %& = \Delta \left[\det \left( - \frac{d^2}{du^2} - \Omega(u)^2  \right) 
1919: %\right]^{-1/2} 
1920: %\earay
1921: Using the result from Eq.(\ref{pref1}), we can write
1922: \baray
1923: \label{ratio}
1924: \frac{\left[\det \left( - \frac{d^2}{du^2} - \Omega(u)^2  \right) 
1925: \right]}{\left[\det \left( - \frac{d^2}{du^2} - U(u)  \right) 
1926: \right]} = \frac{g(u_i) g(u_f) \int_{u_i}^{u_f}\frac{ds}
1927: {g(s)^2}}{G(u_i) G(u_f) \int_{u_i}^{u_f}\frac{ds}
1928: {G(s)^2}}
1929: \earay
1930: where $G(u)$, analogous to $g(u)$, is a solution of the equation
1931: of motion with an arbitrary $U(u)$. And, like $g(u)$, 
1932: $G(u)$ also does not vanish at $u_i$. Now, let $g^0$ denote the
1933: unique solution to $ { \partial_{u}^2 + \Omega(u)^2}g(u) = 0 $
1934: which satisfies the boundary conditions:
1935: \baray
1936: \label{go}
1937: & g^0(u_i) = 0; \quad  & \frac{d}{du}g^0(u_i) = 1
1938: \earay
1939: Let $g^1$ represent the solution that satisfies
1940: \baray
1941: & g^1(u_i)= 1 ;\quad  & \frac{d}{du}g^1(u_i) = 0
1942: \earay
1943: In Eq.(\ref{ratio}), let
1944: \baray
1945: & g(u) = g^0(u) + \epsilon g^1(u) ; \quad  &   G(u) = G^0(u) + \epsilon G^1(u)
1946: \earay
1947: The integral $\int_{u_i}^{u_f}\frac{du}{[g(t)]^2}$ diverges due to
1948: the vanishing of $g^0$ at $u_i$. However, since almost all the contribution
1949: comes from an infinitely small neighborhood of $u_i$ (in the limit
1950: when $\epsilon \to 0$), it follows that it diverges like
1951: \baray
1952: \int_{u_i}^{u_f}\frac{du}{[u - u_i]^2}
1953: \earay
1954: Consequently,
1955: \baray
1956: & {\stackrel{\lim}{\epsilon \to 0}} \quad
1957: {\int_{u_i}^{u_f}\frac{du}{g(u)^2}}=
1958: { \int_{u_i}^{u_f}\frac{du}{G(u)^2}} 
1959: \earay
1960: So Eq.(\ref{ratio}) becomes
1961: \baray
1962: \frac{\left[\det \left( - \frac{d^2}{du^2} - \Omega(u)^2  \right) 
1963: \right]}{\left[\det \left( - \frac{d^2}{du^2} - U(u)  \right) 
1964: \right]} = \frac{g^0(u_f)}{G^0(u_f)}
1965: \earay
1966: Taking $U(u) = 0$ (for a free particle), we get
1967: \baray
1968: \frac{\left[\det \left( - \frac{d^2}{du^2} - \Omega(u)^2  \right) 
1969: \right]}{\left[\det \left( - \frac{d^2}{du^2}\right) 
1970: \right]} = \frac{g^0(u_f)}{u_f - u_i}
1971: \earay
1972: 
1973: Now we can calculate the propagator,
1974: \baray
1975: \label{final}
1976: & K(x_f, T| x_i, 0)  \equiv K(x_f, u_f | x_i, u_i)  = K(0, u_f| 0, u_i)
1977: \exp \left( i S[x_{cl}] \right)  \nonumber \\
1978: & = \left(\frac{\det \left( - \frac{d^2}{du^2} - \Omega(u)^2  \right)} 
1979: {\det \left( - \frac{d^2}{du^2}\right)}\right)^{-1/2} 
1980: K_{0}(0, u_f| 0, u_i) \exp \left( i S[x_{cl}] \right)
1981:  \nonumber \\
1982: & = \sqrt{\frac{1}{2 \pi i g^0(u_f)}} \exp \left( i S[x_{cl}] \right) 
1983: \earay
1984: 
1985: One can switch back to the $t$ variable now using Eq.(\ref{defnu}). 
1986: $S[x_{cl}]$ is the action for the solution $x_{cl}$ of the classical
1987: equation of motion satisfying the boundary conditions $x(u_i) = x_i$
1988: and $x(u_f) = x_f$.
1989: 
1990: Now we specialize this calculation of the path integral for an oscillator
1991: with varying mass and frequency to our case. Eq.(\ref{class_eqn}) generally
1992: does not lend itself to simple analysis. However, since we expect the 
1993: higher modes (with large values of $n$) to be the major contributors
1994: to the suppression of the tunnelling, one can use the WKB analysis
1995: to find approximate solutions. This is because, firstly, Eq.(\ref{class_eqn}) 
1996: has the form of a Schrodinger equation for an
1997: arbitrary potential $\Omega(u)^2 $, and, secondly, the variation
1998: in $\Omega(u)^2 x_{cl}(u)$ for higher modes is slow enough. Typically, for
1999: the $n$th mode, $\Omega^2 \sim n^2$ and $\dot{\Omega} \sim n$. 
2000: So $\dot{\Omega} << \Omega^2$.
2001: Using the WKB approximation, we get:
2002: \baray
2003: x_{cl}(u) = \frac{1}{\sqrt{\Omega}} \exp \left(\pm i
2004: \int_{u_i}^{u_f}du \Omega(u)  \right)
2005: \earay
2006: 
2007: Note that in the Euclidean case, $u \to -i u$ (or, equivalently,
2008: $t \to -i \tau$). This is what we have done in Eq.(\ref{class_soln}, 
2009: \ref{gensol}) to get the solution for the tensor mode $t_{cl}(\tau)$. Using these
2010: classical solutions we have calculated the classical action in
2011: Eq.(\ref{S_E}). 
2012: %In Eq.(\ref{pre_fac}) we have calculated the prefactor. 
2013: As explained in Eq.(\ref{final}), the prefactor is
2014: proportional to $({2 \pi i g^0(u_f)})^{-1/2}$  where
2015: $g^0(u)$ satisfies the boundary conditions given in Eq.(\ref{go}).
2016: Following Eq.(\ref{gensol}) it is clear that for tensor modes,
2017: the two linearly independent solutions to the classical equation of motion 
2018: are $\frac{1}{\sqrt{\omega_n}} \exp \left(\pm \int^u du' \omega_n(u') \right)$
2019: . And the following linear combination satisfies the required boundary 
2020: conditions
2021: \ba
2022: g^0(u_f) = \frac{1}{\sqrt{2\omega_n(u_i)\omega_n(u_f)}}\left( \exp(D_n) 
2023: - \exp(-D_n) \right).
2024: \ea
2025: as this choice of $g^0(u)$ vanishes at $u = u_i$ where $D_n = 0$.
2026: $\frac{dg^0(u)}{du} = 1$ at $u = u_i$.
2027: This yields the result in Eq.(\ref{pre_fac}).
2028: 
2029: 
2030: 
2031: 
2032: \vspace{0.5cm}
2033: 
2034: 
2035: 
2036: \begin{thebibliography}{10}
2037: 
2038: \bibitem{Vilenkin:1982de}
2039: A.~Vilenkin, {\it Creation of universes from nothing},  {\em Phys. Lett.} {\bf
2040:   B117} (1982) 25; 
2041: %\bibitem{Vilenkin:1983xq}
2042: %A.~Vilenkin, 
2043: {\it The birth of inflationary universes},  {\em Phys. Rev.} {\bf
2044:   D27} (1983) 2848.
2045:   
2046: \bibitem{Hartle:1983ai}
2047: J.~B. Hartle and S.~W. Hawking, {\it Wave function of the universe},  {\em
2048:   Phys. Rev.} {\bf D28} (1983) 2960; \\
2049:  S.~W. Hawking, {\it The Quantum State of the Universe}, {\em Nucl. Phys.} {\bf B239} (1984) 257.
2050:  
2051: %\cite{Firouzjahi:2004mx}
2052: \bibitem{Firouzjahi:2004mx}
2053: H.~Firouzjahi, S.~Sarangi and S.-H.~H.~Tye,
2054: {\it Spontaneous creation of inflationary universes and the cosmic landscape},
2055: JHEP {\bf 0409}, 060 (2004), hep-th/0406107.
2056: %%CITATION = HEP-TH 0406107;%%
2057: 
2058: 
2059: %\cite{Linde:1998gs}
2060: \bibitem{Linde:1998gs}
2061: A.~D.~Linde,
2062: {\it Quantum creation of an open inflationary universe},
2063: {\em Phys. Rev. D} {\bf 58}, 083514 (1998),
2064: gr-qc/9802038; \\
2065: %\cite{Vilenkin:1998rp}
2066: A.~Vilenkin,
2067: {\it The quantum cosmology debate},
2068: gr-qc/9812027.
2069: %%CITATION = GR-QC 9812027;%%
2070: 
2071: 
2072: \bibitem{Gibbons:1978ac}
2073: G.~W. Gibbons, S.~W. Hawking, and M.~J. Perry, {\it Path integrals and the
2074:   indefiniteness of the gravitational action},  {\em Nucl. Phys.} {\bf B138}
2075:   (1978) 141; \\
2076: %\cite{Gibbons:1978ji}
2077: %\bibitem{Gibbons:1978ji}
2078: G.~W.~Gibbons and M.~J.~Perry,
2079: {\it Quantizing Gravitational Instantons},
2080: Nucl.\ Phys.\ B {\bf 146}, 90 (1978).
2081: %%CITATION = NUPHA,B146,90;%%
2082: 
2083: %\cite{Brown:1987dd}
2084: \bibitem{Brown:1987dd}
2085: J.~D.~Brown and C.~Teitelboim,
2086: {\it Dynamical Neutralization Of The Cosmological Constant},
2087: {\em Phys. Lett. B} {\bf 195},(1987) 177; \\
2088: %\cite{Bousso:2000xa}
2089: %\bibitem{Bousso:2000xa}
2090:   R.~Bousso and J.~Polchinski,
2091:   {\it Quantization of four-form fluxes and dynamical neutralization of the
2092:   cosmological constant},
2093:   JHEP {\bf 0006}, 006 (2000), hep-th/0004134.
2094: 
2095: %\cite{Coleman:1988tj}
2096: \bibitem{Coleman:1988tj}
2097:  S.~W. Hawking, {\it The Cosmological Constant is Probably Zero},
2098:   {\em Phys. Lett.} {\bf B134} (1984) 403; \\ 
2099:   S.~R.~Coleman,
2100:   {\it Why There Is Nothing Rather Than Something: A Theory Of The Cosmological
2101:   Constant},
2102:   Nucl.\ Phys.\ B {\bf 310}, 643 (1988); \\
2103: %\cite{Weinberg:1988cp}
2104: %\bibitem{Weinberg:1988cp}
2105: S.~Weinberg,
2106: {\it The Cosmological Constant Problem},
2107: {\em Rev. Mod. Phys.}  {\bf 61}, (1989) 1; \\
2108: %\cite{Strominger:1988yt}
2109: %\cite{Giddings:1987cg}
2110: %\bibitem{Giddings:1987cg}
2111:   S.~B.~Giddings and A.~Strominger,
2112:   {\it Axion Induced Topology Change In Quantum Gravity And String Theory},
2113:   Nucl.\ Phys.\ B {\bf 306}, 890 (1988); \\
2114:   %%CITATION = NUPHA,B306,890;%%
2115: %\bibitem{Strominger:1988yt}
2116: %A.~Strominger,
2117: %{\it A Lorentzian Analysis Of The Cosmological Constant Problem},
2118: %{\em Nucl. Phys. B} {\bf 319},(1989) 722; \\
2119: %\cite{Klebanov:1988eh}
2120: %\bibitem{Klebanov:1988eh}
2121: I.~R.~Klebanov, L.~Susskind and T.~Banks,
2122: {\it Wormholes And The Cosmological Constant},
2123: {\em Nucl. Phys. B} {\bf 317}, (1989) 665; \\
2124: %\bibitem{Fischler:1990se}
2125: W.~Fischler, D.~Morgan, and J.~Polchinski, {\it Quantum nucleation of false
2126:   vacuum bubbles},  {\em Phys. Rev.} {\bf D41} (1990) 2638. 
2127:   %where earlier references on the topic can be found.
2128: 
2129: \bibitem{Linde:1984mx}
2130: A.~D. Linde, {\it Quantum creation of the inflationary universe},  {\em Nuovo
2131:   Cim. Lett.} {\bf 39} (1984) 401; \\
2132: %\bibitem{Vilenkin:1984wp}
2133: A.~Vilenkin, {\it Quantum Creation Of Universes}, Phys.\ Rev.\ D {\bf 30}, 509 (1984).
2134:  
2135: %\cite{Rubakov:1984bh}\cite{Lavrelashvili:1985vn}
2136: \bibitem{Rubakov:1984bh}
2137: V.~A.~Rubakov,
2138: {\it Quantum Mechanics In The Tunneling Universe},
2139: Phys.\ Lett.\ B {\bf 148}, 280 (1984); \\
2140: %\cite{Lavrelashvili:1985vn}
2141: %\bibitem{Lavrelashvili:1985vn}
2142: G.~V.~Lavrelashvili, V.~A.~Rubakov and P.~G.~Tinyakov,
2143: {\it Tunneling Transitions With Gravitation: Breaking Of The Quasiclassical
2144: Approximation},
2145: Phys.\ Lett.\ B {\bf 161}, 280 (1985).
2146: 
2147: %\cite{Hawking:1976ja}
2148: \bibitem{Hawking:1976ja}
2149:   S.~W.~Hawking,
2150:   %``Zeta Function Regularization Of Path Integrals In Curved Space-Time,''
2151:   Commun.\ Math.\ Phys.\  {\bf 55}, 133 (1977).
2152:   %%CITATION = CMPHA,55,133;%%
2153: 
2154: %\cite{Gibbons:1978ji}
2155: \bibitem{Gibbons:1978ji}
2156:   G.~W.~Gibbons and M.~J.~Perry,
2157:   %``Quantizing Gravitational Instantons,''
2158:   Nucl.\ Phys.\ B {\bf 146}, 90 (1978).
2159:   %%CITATION = NUPHA,B146,90;%%
2160:   
2161: %\cite{Ooguri:2005vr}
2162: \bibitem{Ooguri:2005vr}
2163: H.~Ooguri, C.~Vafa and E.~Verlinde,
2164: {\it Hartle-Hawking wave-function for flux compactifications},
2165: hep-th/0502211.
2166: %%CITATION = HEP-TH 0502211;%%
2167: 
2168: %\cite{Coleman:1977py}
2169: \bibitem{Coleman:1977py}
2170: S.~R.~Coleman,
2171: {\it The Fate Of The False Vacuum. 1. Semiclassical Theory},
2172: Phys.\ Rev.\ D {\bf 15}, 2929 (1977)
2173: [Erratum-ibid.\ D {\bf 16}, 1248 (1977)].
2174: %%CITATION = PHRVA,D15,2929;%%
2175: 
2176: \bibitem{Sethna:1981dr}
2177: J.~P.~Sethna, {\it Decay rates of tunneling centers 
2178: coupled to phonons : An instanton approach},
2179: {\em Phys. Rev.} {\bf B25} (1981) 5050; \\
2180: %\bibitem{sethna:1981ds}
2181: {\it Phonon coupling in tunneling systems 
2182: at zero temperature : An instanton approach},
2183: {\em Phys. Rev.} {\bf B24} (1981) 698.
2184: 
2185: %\cite{Caldeira:1982uj}
2186: \bibitem{Caldeira:1982uj}
2187: A.~O.~Caldeira and A.~J.~Leggett,
2188: {\it Quantum Tunneling In A Dissipative System},
2189: Annals Phys.\  {\bf 149}, 374 (1983).
2190: %%CITATION = APNYA,149,374;%%
2191: 
2192: %\bibitem{Dewitt:1967yk}
2193: %B.~S.~Dewitt, {\it Quantum Theory Of Gravity. 1. The Canonical Theory},
2194: %{\em Phys. Rev.}  {\bf 160} (1967) 1113.
2195: 
2196: \bibitem{zeh}
2197: D.~Giulini, E.~Joos, C.~Kiefer, J.~Kupsch, I.-O.~Stamatescu and H.~D.~Zeh,
2198: {\it Decoherence and the Appearance of a Classical World in Quantum Theory}
2199: {Berlin; New York: Springer, 1996};
2200: W.~H.~Zurek, {\it Decoherence and the transition from quantum to classical},
2201:  Physics Today,  {\bf 44}, {\it 10}, 36 (1991);
2202:  M.~Gell-mann and J.B.~Hartle, {\it Classical Equations for Quantum Systems},
2203:  Phys.\ Rev.\ D {\bf 47}, 3345 (1993).
2204: 
2205: \bibitem{Guth:1981zm}
2206: A.~H.~Guth, {\it The inflationary universe: A possible solution to the horizon
2207:   and flatness problems},  {\em Phys. Rev.} {\bf D23} (1981) 347; \\
2208: %\bibitem{Linde:1982mu}
2209: A.~D.~Linde, {\it A new inflationary universe scenario: A possible solution of
2210:   the horizon, flatness, homogeneity, isotropy and primordial monopole
2211:   problems},  {\em Phys. Lett.} {\bf B108} (1982) 389; \\
2212: %\bibitem{Albrecht:1982wi}
2213: A.~Albrecht and P.~J.~Steinhardt, {\it Cosmology for grand unified theories
2214:   with radiatively induced symmetry breaking},  {\em Phys. Rev. Lett.} {\bf 48}
2215:   (1982) 1220.
2216: 
2217: %\bibitem{Tryon:1973}
2218: %E.~P. Tryon, {\it Is the universe a vacuum fluctuation ?},  {\em Nature} {\bf   246} (1973) 396.
2219: 
2220: \bibitem{Dvali:1998pa}
2221: G.~R. Dvali and S.-H.~H. Tye, {\it Brane inflation},  {\em Phys. Lett.} {\bf
2222:   B450} (1999) 72, hep-ph/9812483.
2223: 
2224: \bibitem{Kachru:2003sx}
2225: S.~Kachru, R.~Kallosh, A.~Linde, J.~Maldacena, L.~McAllister, and S.~P.
2226:   Trivedi, {\it Towards inflation in string theory},  {\em JCAP} {\bf 0310}
2227:   (2003) 013, hep-th/0308055.
2228: 
2229: \bibitem{Halliwell:1985eu}
2230: J.~J. Halliwell and S.~W. Hawking, {\it The origin of structure in the
2231:   universe},  {\em Phys. Rev.} {\bf D31} (1985) 1777.
2232: 
2233: \bibitem{Kiefer:1987ft}
2234: C.~Kiefer, {\it Continuous measurement of minisuperspace variables by higher
2235:   multipoles},  {\em Class. Quant. Grav.} {\bf 4} (1987) 1369.
2236: 
2237: \bibitem{Kiefer:1989ud}
2238: C.~Kiefer, {\it Continuous measurement of intrinsic time by fermions},  {\em
2239:   Class. Quant. Grav.} {\bf 6} (1989) 561.
2240: 
2241: \bibitem{Halliwell:1989vw}
2242: J.~J. Halliwell, {\it Decoherence in quantum cosmology},  {\em Phys. Rev.} {\bf
2243:   D39} (1989) 2912.
2244: 
2245: \bibitem{Kiefer:1992cn}
2246: C.~Kiefer, {\it Decoherence in quantum electrodynamics and quantum gravity},
2247:   {\em Phys. Rev.} {\bf D46} (1992) 1658;\\
2248: A.~O.~Barvinsky, A.~Yu.~Kamenshchik,
2249: C.~Kiefer and I.~V.~Mishakov, {\it Decoherence in quantum cosmology at the onset of inflation}, {\em Nucl. Phys.} {\bf B 551} (1999) 374-396,
2250: gr-qc/9812043
2251: 
2252: 
2253: 
2254: %\cite{Gerlach:1978gy}
2255: \bibitem{Gerlach:1978gy}
2256: U.~H.~Gerlach and U.~K.~Sengupta,
2257: {\it Homogeneous Collapsing Star: Tensor And Vector Harmonics 
2258: For Matter And Field Asymmetries},
2259: {\em Phys.Rev.} {\bf D18} (1978) 1773.
2260: %%CITATION = PHRVA,D18,1773;%%
2261: 
2262: %\cite{Barvinsky:1993nf}
2263: \bibitem{Barvinsky:1993nf}
2264:   A.~O.~Barvinsky,
2265:  {\it Tunneling geometries. 2. Reduction methods for functional determinants},
2266:   Phys.\ Rev.\ D {\bf 50}, 5115 (1994),
2267:   gr-qc/9311023.
2268: 
2269: \bibitem{Harrison:1967ab}
2270: E.~R.~Harrison,
2271: {\it Classification of uniform cosmological models}, {\em Mon. 
2272: Not. R. Astron. Soc.} {\bf 137} (1967) 69.
2273: 
2274: %\cite{Arkani-Hamed:1998rs}
2275: \bibitem{Arkani-Hamed:1998rs}
2276:   N.~Arkani-Hamed, S.~Dimopoulos and G.~R.~Dvali,
2277:   {\it The hierarchy problem and new dimensions at a millimeter},
2278:   Phys.\ Lett.\ B {\bf 429}, 263 (1998)
2279:   hep-ph/9803315.
2280:   %%CITATION = HEP-PH 9803315;%%
2281:   
2282:   %\cite{Hsu:2004jt}
2283: \bibitem{Hsu:2004jt}
2284:   S.~Hsu and A.~Zee,
2285:  {\it A speculative relation between the cosmological constant and the Planck mass},
2286:   hep-th/0406142; \\
2287:   %%CITATION = HEP-TH 0406142;%%
2288:  %\cite{Kobakhidze:2004gm}
2289: %\bibitem{Kobakhidze:2004gm}
2290:   A.~Kobakhidze and L.~Mersini-Houghton,
2291:   {\it Birth of the universe from the landscape of string theory},
2292: hep-th/0410213; \\
2293:   %%CITATION = HEP-TH 0410213;%%
2294:   %\cite{Mersini-Houghton:2005im}
2295: %\bibitem{Mersini-Houghton:2005im}
2296:   L.~Mersini-Houghton,
2297:   {\it Can we predict Lambda for the non-SUSY sector of the landscape?},
2298:   hep-th/0504026.
2299:   %%CITATION = HEP-TH 0504026;%%
2300:  
2301: \bibitem{Kachru:2003aw}
2302: S.~Kachru, R.~Kallosh, A.~Linde, and S.~P. Trivedi, {\it De sitter vacua in
2303:   string theory},  {\em Phys. Rev.} {\bf D68} (2003) 046005,
2304:      hep-th/0301240.
2305: 
2306: \bibitem{Douglas:2003um}
2307: M.~R. Douglas, {\it The statistics of string / m theory vacua},  {\em JHEP}
2308:   {\bf 05} (2003) 046, 
2309:     hep-th/0303194.
2310: 
2311: %\cite{Barvinsky:1992dz}
2312: \bibitem{Barvinsky:1992dz}
2313: A.~O.~Barvinsky, A.~Y.~Kamenshchik and I.~P.~Karmazin,
2314: {\it One loop quantum cosmology: Zeta function technique for the Hartle-Hawking
2315: wave function of the universe},
2316: Annals Phys.\  {\bf 219}, 201 (1992).\\
2317: %%CITATION = APNYA,219,201;%%
2318: A.~O.~Barvinsky and A.~Yu.~Kamenshchik, 
2319: {\it Tunnelling Geometries: Analyticity, Unitarity and Instantons in Quantum Cosmology}
2320: , {\em Phys. Rev.} {\bf D 50} (1994) 5093-5114, gr-qc/9311022.
2321: 
2322: 
2323: %\cite{Barvinsky:1996ce}
2324: \bibitem{Barvinsky:1996ce}
2325: A.~O.~Barvinsky, A.~Y.~Kamenshchik and I.~V.~Mishakov,
2326: {\it Quantum origin of the early inflationary universe},
2327: {\em Nucl. Phys. B} {\bf 491}, 387 (1997),
2328: gr-qc/9612004.\\
2329: %%CITATION = GR-QC 9612004;%%
2330: %\cite{Barvinsky:1999qn}
2331:  A.~O.~Barvinsky, A.~Y.~Kamenshchik and C.~Kiefer,
2332: {\it Origin of the inflationary universe}
2333: {\em Mod. Phys. Lett.}  {\bf A 14}, 1083 (1999),
2334:  gr-qc/9905098.
2335:   %%CITATION = GR-QC 9905098;%%
2336:  
2337: %\cite{Felsager:1981iy}
2338: \bibitem{Felsager:1981iy}
2339:   B.~Felsager,
2340:  {\it Geometry, Particles And Fields}, {Springer, 1981}
2341: %\href{http://www.slac.stanford.edu/spires/find/hep/www?irn=998630}{SPIRES entry}
2342: 
2343: \bibitem{Kane:2004ct}
2344:   G.~L.~Kane, M.~J.~Perry and A.~N.~Zytkow,
2345:  {\it An approach to the cosmological constant problem(s)},
2346:   Phys.\ Lett.\ B {\bf 609}, 7 (2005),
2347: hep-ph/0408169.
2348: 
2349: 
2350: 
2351: 
2352: 
2353: \end{thebibliography}
2354: 
2355: 
2356: \end{document}
2357: 
2358: