1: %\documentclass[preprint,aps,preprintnumbers,amsmath,amssymb]{revtex4}
2: \documentclass[preprint,nofootinbib,aps,preprintnumbers,amsmath,amssymb]{revtex4}
3: %documentclass[11pt]{JHEP3}
4: %\documentclass[12pt]{JHEP3}
5: \usepackage{graphicx}% Include figure files
6: \usepackage{dcolumn}% Align table columns on decimal point
7: \usepackage{bm}% bold math
8: \usepackage{natbib}
9:
10: \newcommand{\be}{\begin{equation}}
11: \newcommand{\ee}{\end{equation}}
12: \newcommand{\beq}{\begin{equation}}
13: \newcommand{\eq}{\end{equation}}
14: \newcommand{\bea}{\begin{eqnarray}\displaystyle}
15: \newcommand{\eea}{\end{eqnarray}}
16: \newcommand{\ea}{\end{eqnarray}}
17: \newcommand{\ph}{$\phi$ }
18: \newcommand{\ch}{$\chi$ }
19: \newcommand{\p}{\partial}
20: \newcommand{\s}{\sigma}
21: \newcommand{\al}{\alpha}
22: \newcommand{\half}{\frac{1}{2}}
23: \newcommand{\pp}{p^+}
24: \newcommand{\vol}{V_0^{(7)}}
25:
26: %% Adjust these for your printer:
27: \textwidth=6.3in \textheight=8.2in
28: \oddsidemargin=0.1in \topmargin=.03cm
29: %%%%%%%%%%%%%%%%%%%%%%%% BEGIN DOCUMENT %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
30: \begin{document}
31:
32: \title{Dilaton Dynamics from Production of Tensionless Membranes}
33:
34: \author{Sera Cremonini}
35: \email{sera@het.brown.edu}
36: \affiliation{Physics Department, Brown University\\
37: Providence, RI 02906}
38: \author{Scott Watson}
39: \email{watsongs@physics.utoronto.ca}
40: \affiliation{Physics Department, University of Toronto \\
41: Toronto, ON}
42:
43: \date{\today}
44:
45: \begin{abstract}
46: In this paper we consider classical and quantum corrections to
47: cosmological solutions of $11$D SUGRA coming from dynamics of
48: membrane states. We first consider the supermembrane spectrum
49: following the approach of Russo and Tseytlin for consistent
50: quantization. We calculate the production rate of BPS membrane
51: bound states in a cosmological background and find that such
52: effects are generically suppressed by the Planck scale, as
53: expected. However, for a modified brane spectrum possessing
54: enhanced symmetry, production can be finite and significant. We
55: stress that this effect could not be anticipated given only a
56: knowledge of the low-energy effective theory. Once on-shell,
57: inclusion of these states leads to an attractive force pulling the
58: dilaton towards a fixed point of S-duality, namely $g_s=1$.
59: Although the SUGRA description breaks down in this regime,
60: inclusion of the enhanced states suggests that the center of
61: M-theory moduli space is a dynamical attractor. Morever, our
62: results seem to suggest that string dynamics does indeed favor a
63: vacuum near fixed points of duality.
64: \end{abstract}
65:
66: \maketitle
67: \tableofcontents
68: \newpage
69:
70: \section{Introduction}
71: Low-energy descriptions of string theory generically predict the
72: existence of a large number of scalar fields,
73: or moduli, which are associated with the size and shape of the extra
74: dimensions,
75: as well as the position and orientation of any branes present in the
76: theory.
77: These moduli are of interest for a number of reasons.
78: From a theoretical point of view, the different vacuum expectation
79: values (VEVs) of these
80: fields correspond to different choices for the string vacuum, leading
81: to an indeterminacy of the theory.
82: From a more phenomenological viewpoint, light scalars can have a profound
83: effect
84: on both the early and late universe. If not fixed, these moduli can
85: lead to a period of
86: inflation, modifications of fundamental constants, and violations of
87: the equivalence principle.
88: If fixed, one must worry about the mass scales involved and the
89: effects of the resulting relic density on cosmological
90: observations, e.g. on the cosmic microwave background
91: or Big Bang nucleosynthesis.
92:
93: Recently it has been argued \cite{Vafa:2005ui} (see also
94: \cite{Arkani-Hamed:2006dz}) that, by taking low-energy
95: supergravity (SUGRA) as the effective description for string theory, we
96: may
97: miss certain crucial aspects of the underlying theory.
98: As an example, when attempting to understand the vacuum
99: structure of the theory one
100: should remember that moduli space must be of finite size in order to
101: have a realistic theory of gravity (finite $G_N$).
102: Other crucial aspects that we wish to address in this paper are the
103: role of dualities and the importance of dynamics.
104: That is, as background fields evolve the effective
105: mass of heavy states that are outside the realm of the low-energy
106: theory can change.
107: In fact, near points of enhanced symmetry (ESPs), frequently associated
108: with dualities, these additional states can become massless and
109: therefore play a vital role in the low-energy theory.
110: It is important that these states lie beyond the naive low-energy
111: SUGRA description and, without an underlying knowledge of the
112: fundamental theory,
113: effects associated with these additional degrees of freedom (such as
114: particle creation and radiative corrections) would be missed.
115: Indeed, recent work suggests that including these effects can have
116: important and interesting effects both in string theory and cosmology
117: \cite{Silverstein:2003hf,Kofman:2004yc,Watson:2004aq,Alishahiha:2004eh}.
118:
119: In this paper we will consider cosmological solutions of $11$D SUGRA
120: and ask what corrections, if any, result from the dynamics of the
121: moduli (i.e., dilaton and radii). Focusing on the case of particle
122: production, we show that production of BPS membrane bound states,
123: whose mass depends on the evolution of the moduli, is generally
124: suppressed by the Planck scale. This is expected and assures us of
125: the validity of the low-energy effective theory. However, in the
126: special case of membranes that become tensionless near fixed points of
127: duality, we find that significant production can occur. We are unaware of an explicit construction of such states, however we anticipate their existence given ubiquitous examples in the
128: lower-dimensional string theory case.
129:
130: We find that by including enhanced states in the low-energy theory, the
131: center of the M-theory moduli space becomes an attractor, and the
132: evolution of the moduli tends toward fixed points of duality. In the
133: case of the dilaton, which can be taken to be $R_{11}$ in the $11$D
134: theory, we find that the evolution leads to the region near $g_s=1$,
135: i.e. strong coupling.
136: Naively trusting the
137: equations of motion in this region, we find that all moduli will in
138: fact become trapped and we recover a three dimensional radiation
139: dominated universe at late times. Although we expect additional
140: corrections to arise near the regime of strong coupling, it is
141: intriguing that our results seem to provide an
142: explicit example of the belief (\cite{Brustein:2002xf} and
143: \cite{Dine:1998qr}) that a string vacuum consistent with gravity and
144: particle phenomenology should lie near fixed points of duailty.
145:
146:
147: The paper is organized as follows. In Section {\ref{MthStates} we
148: review a scheme developed by Russo and Tseytlin for quantizing the
149: supermembrane in special limits. This will allow us to obtain the BPS
150: spectrum of membrane bound states, which we will utilize throughout
151: the paper.
152: In Section \ref{ClTrapping} we briefly discuss cosmology with these
153: states and show that they can lead to stabilization, but we also point out
154: that
155: such considerations seem inconsistent, since they lie outside the scope
156: of the effective theory. In Section \ref{QmTrapping} we consider the
157: possibility of producing membrane bound states within a cosmological
158: background of $11$D SUGRA. After reviewing the basic formalism for
159: calculating production, we find an expression for the production rate
160: and find that membrane bound states suffer exponential suppression, as
161: expected. However, we also find that, for a special class of states
162: possessing enhanced symmetry, significant production can result.
163: In Section \ref{cosmo} we consider the modified effective theory in
164: the presence of the produced states, and determine the new evolution
165: of the moduli. We find an attractor behavior towards the center of
166: moduli space, and trusting the equations in the center region leads to
167: trapping of the moduli. We conclude with a discussion of the limitations of our approach and future work in progress.
168:
169: \section{M-theory and the Supermembrane \label{MthStates}}
170:
171: In this section we review progress that has been made in understanding M-theory
172: as a fundamental theory of membranes.
173: The goal will be to gain an understanding of the spectrum of the fundamental theory
174: beyond the low-energy effective description of $11$D SUGRA.
175: We will be concerned with finding membrane states whose mass depends explicitly
176: on the low-energy moduli, namely the dilaton and radii of the extra dimensions.
177:
178: We begin by reviewing the work of \cite{Russo:1996if}, where
179: the authors were able to clarify the $11$D origin of many
180: lower-dimensional solutions, as well as identify states of the supermembrane
181: with known string theory configurations.
182: Much effort has been devoted to understanding composite BPS
183: configurations of branes in $11$D.
184: We will concentrate in particular on non-threshold (non-zero
185: binding energy) bound states, the canonical example of which (in string theory) is
186: the $(p,q)$ string, a bound
187: state of a NS-NS string and a R-R string in type IIB theory.
188:
189: In \cite{Schwarz:1995du,Schwarz:1996bh}, it was
190: suggested that the $(p,q)$ string states should be related to the
191: BPS states of a wrapped M$2$-brane in $11$D. This was supported by a
192: comparison of the {\it zero-mode} contributions of the membrane to the
193: string mass spectra, where agreement was found. However, matching at
194: the oscillator level was not shown, due to the intrinsic
195: difficulties associated with the quantization of the supermembrane.
196: This difficulty was overcome and
197: the {\it oscillating} contributions to the spectrum were
198: obtained in \cite{Russo:1996if}, where supermembrane
199: quantization was achieved in a specific limit in which the theory
200: dramatically simplifies.
201:
202: Some of the M-theory
203: solutions studied in \cite{Russo:1996if} can have interesting
204: cosmological consequences, as will become apparent in the next section.
205: Specifically, we are interested in certain M-theory states
206: which have a classical supergravity description in terms of
207: a bound state of an M$2$-brane wrapping a $T^2$
208: and of a gravitational wave propagating along one of the torus directions.
209: We will start by presenting the 11D SUGRA solution for such configurations, and show
210: that in a certain limit it can be reduced to the type IIB $(p,q)$ string.
211: Since we are primarily interested in obtaining the mass spectrum of these states, we will
212: then present their interpretation in supermembrane theory.
213: Specifically, we will outline the derivation of the spectrum of such
214: solutions as given in \cite{Russo:1996if}, where it was first confirmed that
215: the mass of an oscillating membrane does indeed agree with that of the $(p,q)$ string.
216: Finally, at the end of this section we attempt to motivate a new class of states possessing enhanced symmetry. We will save an explicit construction of such states for future work, here only comparing their spectrum to the know BPS states of the supermembrane. We will see in future sections that such states are the only possibility for membrane production in a low-energy theory.
217:
218: \subsection{D=11 SUGRA Solutions}
219:
220: We present a concise introduction to the 11D origin of certain
221: non-marginal BPS configurations of type II string theories.
222: We consider the 11-dimensional space $M^9 \times T^2$.
223: The torus coordinates are labeled by $(y_1,y_2)$ and have periods
224: $(2\pi R_9,2\pi R_{11})$, while the spatial coordinates of $M^9$ are denoted by $x_i$, $i=1,\ldots,8$.
225: We are ultimately interested in considering a bound state of an M$2$ brane,
226: which wraps the $T^2$, and of a gravitational wave propagating in
227: the $y_2$ direction. We will briefly show how to construct such states
228: from simpler solutions.
229:
230: The 11D solution representing a gravitational wave moving along $y_1$ is given by
231: \bea
232: ds^2 &=& -dt^2 + dy_1^2 + dy_2^2 + W (d t+dy_1)^2+ dx_i dx_i \, ,\nonumber \\
233: C_{\mu \nu \rho} &=& 0, \; \; \; W=\frac{\tilde{Q}}{r^6}, \; \; \;
234: r^2=x_i x_i.
235: \ea
236: Since $y_1$ is periodic, the charge $\tilde{Q}$ must be quantized
237: in units of $1/R_9$; with the correct normalization factor, the
238: charge becomes $\tilde{Q}=c_0 \frac{l}{R_9}$.
239: This solution is found to preserve $1/2$ of the supersymmetries.
240: The other basic 11D object that we will need is the 2-brane, with
241: solution \cite{Russo:1996if},
242: \bea
243: ds^2 &=& H_2^{-\frac{2}{3}} \, \left[-dt^2 + dy_1^2 + dy_2^2 \right] + H_2^{\frac{1}{3}} dx_i
244: dx_i \, , \nonumber \\
245: C_3 &=& H_2^{-1} dt \wedge dy_1 \wedge dy_2, \;\;\;
246: H_2=1+\frac{Q}{r^6}.
247: \ea
248: The charge in this case is given by $Q=c_0 \frac{w_0 R_9}{\alpha'}$, where
249: $w_0$ denotes winding around the target torus (we will define
250: $w_0$ more precisely in the next section).
251:
252:
253: Combining the 2-brane solution with a gravitational wave moving in
254: an {\it arbitrary} direction gives
255: \bea
256: ds^2 &=& H_2^{-\frac{2}{3}} \, \bigl[-dt^2 + dy_1^2 + dy_2^2 + W (d t-dz_1)^2\bigr]+ H_2^{\frac{1}{3}} dx_i
257: dx_i\, , \nonumber \\
258: C_3 &=& H_2^{-1} dt \wedge dy_1 \wedge dy_2, \; \; \; z_1=y_1
259: \cos{\theta}+y_2\sin{\theta}\, ,
260: \ea
261: where the gravitational wave charge is modified
262: in the following way:
263: \bea
264: W&=&\frac{\tilde{Q}_q}{r^6}, \; \; \; \tilde{Q}_q=c_0\sqrt{\frac{l_9^2}{R_9^2}+\frac{l_{11}^2}{R_{11}^2}}
265: =\tilde{Q}\sqrt{q_1^2+q_2^2 \tau_2^2},\\
266: \text{with}&& \; \; \; \cos{\theta}=\frac{q_1}{\sqrt{q_1^2+q_2^2 \tau_2^2}}, \; \; \; \; \;
267: \tau_2=\frac{R_9}{R_{11}}.
268: \ea
269: After appropriate dimensional reductions and applications of dualities,
270: this solution can be identified with a boosted BPS bound state of
271: an $F1$ string and $D0$ brane in Type IIA string theory, and with
272: a boosted $(p,q)$ string in Type IIB.
273:
274: We are specifically interested in the 2-brane + wave solution with
275: a boost along $y_2$ only (with $\cos{\theta}=0$ and $\sin{\theta}=1$),
276: giving \cite{Russo:1996if},
277: \bea
278: ds^2 &=& H_2^{-\frac{2}{3}} \, \bigl[-dt^2 + dy_1^2 + dy_2^2 + W (d t-dy_2)^2\bigr]+ H_2^{\frac{1}{3}} dx_i
279: dx_i, \nonumber \\
280: C_3 &=& H_2^{-1} dt \wedge dy_1 \wedge dy_2,
281: \ea
282: where
283: \bea
284: H_2&=&1+\frac{Q}{r^6}, \; \; \; r^2=x_i x_i, \; \; \; Q= (4\pi^2 T_3 c_0 w_0 R_9)\, R_{11}, \\
285: W&=& \frac{\tilde{Q}}{r^6}, \; \; \; \tilde{Q}=c_0
286: \frac{l_{11}}{R_{11}},
287: \ea
288: and we used the definition of $\alpha'$. This state is BPS, preserving $1/4$ of the supersymmetries.
289:
290: To obtain the mass spectrum of the M-theory membrane that
291: corresponds to this classical supergravity solution, we will make use of
292: supermembrane theory.
293: Before proceeding, we would like to stress the advantage of studying
294: BPS solutions. The BPS condition guarantees that the above classical
295: SUGRA solutions will exhibit some of the features of the full
296: quantum (string) theory, since these states are protected from quantum corrections.
297: That is, if we construct these states in a limit of the theory that is well understood, we can then
298: extrapolate them to regimes that are less understood,
299: giving us a partial knowledge of the spectrum of the theory.
300:
301: \subsection{Supermembrane Mass Spectrum}
302:
303: The original studies of the physical spectrum of wrapped membranes of toroidal topology
304: have been performed in \cite{Bergshoeff:1987qx,Duff:1987cs}.
305: Quantization of the supermembrane is highly non-trivial, and one
306: might even wonder whether a consistent quantum theory of the
307: supermembrane can be defined. Addressing the {\it oscillating}
308: membrane is particularly difficult, since it involves dealing with
309: a highly non-linear interacting theory.
310: In the work of \cite{Russo:1996if}, however, quantization was
311: achieved in an appropriate limit in which the interacting terms
312: dropped out, and the theory could be solved.
313: Next, we will outline the arguments of
314: \cite{Russo:1996if} leading to the mass spectrum of the oscillating
315: membrane.
316:
317: We consider a membrane on $\mathcal{M}^9 \times T^2$. The compact
318: directions are labelled by $X^9$ and $X^{11}$, with radii $R_9,
319: R_{11}$.
320: Wrapping the membrane around the toroidal directions
321: gives
322: \beq
323: \label{X1X2}
324: X^9(\sigma,\rho)= w_9 \, R_9\, \sigma +
325: \tilde{X}^9(\sigma,\rho)\, , \;\;\;\;\;\;\;\;\;
326: X^{11}(\sigma,\rho)= w_{11} \, R_{11}\, \rho +
327: \tilde{X}^{11}(\sigma,\rho)\, ,
328: \eq
329: where the (single-valued) functions
330: $\tilde{X}^9(\sigma,\rho)$ and $\tilde{X}^{11}(\sigma,\rho)$
331: can be expanded in a complete basis of functions on the torus:
332: \beq
333: \label{X1X2tilde}
334: \tilde{X}^9(\sigma,\rho)=\sqrt{\alpha'}\,\sum_{k,m}X^9_{(k,m)}\,
335: e^{ik\sigma+im\rho} , \;\;\;\;\;\;
336: \tilde{X}^{11}(\sigma,\rho)=\sqrt{\alpha'}\,\sum_{k,m}X^{11}_{(k,m)}\,
337: e^{ik\sigma+im\rho}\, . \eq The constants $\alpha'$ and $T_3$ are
338: defined as \beq \alpha'=\frac{1}{4 \pi^2 R_{11} T_3}, \; \; \; \;
339: \; T_3=\frac{l_P^{-3}}{4\pi^2}.
340: \eq
341: It will turn out to be convenient to define a
342: winding number $w_0$ in the following way,
343: \beq
344: w_0 =
345: \frac{1}{(2\pi R_9) (2\pi R_{11})} \int d\sigma d\rho \,
346: \{X^9,X^{11}\}=w_9 w_{11}, \;\;\;\;\; \{X,Y\}\equiv\p_\sigma X\p_\rho
347: Y-\p_\rho X\p_\sigma Y, \nonumber
348: \eq
349: counting how many times the
350: membrane is wound around the target torus. A membrane with $w_0
351: \neq 0$ is stable for topological reasons \cite{Russo:1996ph}, as
352: will become explicit in the Hamiltonian description. This is a
353: motivation for wrapping the membrane on $\mathcal{M}^9 \times T^2$,
354: and not on $\mathcal{M}^{10} \times S^1$, which would give $w_0=0$.
355:
356: The transverse (single-valued) coordinates $X^i$, $i=1, \ldots ,8$, and their
357: corresponding canonical momenta can also be expanded on the torus:
358: \beq
359: \label{Xi}
360: X^i(\sigma,\rho)=\sqrt{\alpha'}\,\sum_{k,m}X^i_{(k,m)}\,
361: e^{ik\sigma+im\rho}, \; \; \; \; \; \; \; \;
362: P^i(\sigma,\rho)=\frac{1}{4 \pi^2
363: \sqrt{\alpha'}}\,\sum_{k,m}P^i_{(k,m)}\,
364: e^{ik\sigma+im\rho}.
365: \eq
366: The (bosonic) light-cone Hamiltonian for the
367: supermembrane is given by
368: \beq
369: H_B=2\pi^2 \int d\s d\rho \bigl[(P^{\,c})^{2} + \frac{1}{2}T_3^2 (\{X^c,X^d\})^2
370: \bigr],
371: \eq
372: where $c,d=1,\ldots,10$.
373: Here we neglect the fermionic sector, which can however be
374: incorporated.
375: Making use of the expansions (\ref{X1X2}),(\ref{X1X2tilde}),(\ref{Xi}), and
376: separating the contributions from the winding modes, one finds
377: \footnote{For a rigorous and detailed derivation of the Hamiltonian see
378: \cite{Russo:1996if}. Here we are interested in outlining only the
379: major steps needed to obtain the mass spectrum.}
380: \bea
381: H_B&=&H_0+H_{int}, \nonumber \\
382: \label{H0}
383: \alpha' H_0 &=& \frac{R_9^2 w_0^2}{2\alpha'}
384: +\frac{1}{2}\sum_{k,m} [P_{(k,m)}^a P_{(-k,-m)}^a + \omega^2_{k\,m}X_{(k,m)}^a X_{(-k,-m)}^a]\\
385: \alpha' H_{int}&=& \frac{1}{4 g^2} \sum_{m,n,p} \sum_{k,l,l'}
386: (pk'-m'k)(nl-ml')
387: X_{(l',n)}^a X_{(k,p)}^a X_{(k',m')}^b X_{(l,m)}^b + \nonumber \\
388: && + \frac{iw_{11}}{g} \sum_{k,m,n} m k^2 X_{2(0,m)}
389: X_{(k,n)}^iX_{(-k,-n-m)}^i,
390: \ea
391: where
392: \bea
393: a,b &=&1, \ldots, 8,11, \;\;\;\; m'=-m-n-p, \; \; \; \; k'=-k-l-l' \nonumber \\
394: g^2 &\equiv & \frac{R_{11}^2}{\alpha'}=4\pi^2 R_{11}^3 T_3, \; \; \; \; \;
395: \omega_{k m} =\sqrt{w_{11}^2 k^2+w_9^2m^2\tau_2^2}, \; \; \;
396: \tau_2=\frac{R_9}{R_{11}}.
397: \ea
398: This is clearly a highly non-linear interacting theory.
399: However, notice that the interacting terms are of order ${\cal O}(\frac{1}{g})$
400: and ${\cal O}(\frac{1}{g^2})$. In the large $g$ limit, such terms
401: are negligible, and can be dropped.
402: Thus, in the limit $g\rightarrow\infty$, given by
403: \beq
404: \label{limit}
405: R_9,R_{11} \rightarrow \infty, \; \; T_3
406: \rightarrow 0 \; \; \; \; (\text{holding} \;\; \alpha' \;\; \text{and} \;\; \tau_2 \;\;
407: \text{fixed})
408: \eq
409: the Hamiltonian reduces to a system of decoupled
410: harmonic oscillators, and the theory can be solved exactly.
411: In particular, we can quantize the system and determine the mass
412: spectrum.
413: At this point it is important to note that since the states we are
414: interested in considering are BPS, their mass is exact, and can be trusted for all
415: radii. Thus, even though the theory was solved for the special
416: limit (\ref{limit}), such states can be studied more generically,
417: a valuable consequence of the BPS condition.
418:
419: We should also point out that as long as $w_0=w_9 \, w_{11}\neq0$,
420: the spectrum of the Hamiltonian is discrete. If one was
421: considering $R^{10}\times S^1$, there would be flat directions in
422: the Hamiltonian, causing the membrane to be unstable, and the
423: spectrum would be continuous.
424:
425: After having dropped $H_{int}$ one can proceed to the quantization
426: of $H_0$. The fields can be expanded in terms of creation and annihilation
427: operators ($a=1,\ldots,8,11 $),
428: \bea
429: X^a_{(k,m)}&=&\frac{i}{\sqrt{2}\,\omega_{(k,m)}} \,
430: [\alpha^a_{(k,m)}+\tilde{\alpha}^a_{(-k,-m)}]\, , \;\;\;\;\;\;
431: P^{\,a}_{(k,m)}=\frac{1}{\sqrt{2}} \,
432: [\alpha^a_{(k,m)}-\tilde{\alpha}^a_{(-k,-m)}]\, , \nonumber \\
433: (X^a_{(k,m)})^\dag &=& X^a_{(-k,-m)}\, , \; \; \; \;(P^{\,a}_{(k,m)})^\dag =P^{\,a}_{(-k,-m)},
434: \;\;\;\;\omega_{(k,m)}=\text{sign}(k)\, \omega_{km}\, .
435: \ea
436: Canonical commutation relations give
437: \beq
438: \left[ X^a_{(k,m)},P^b_{(k',m')}\right] = i
439: \delta_{k+k'}\delta_{m+m'}\delta^{ab},\;\;\;\;
440: \left[ \alpha^a_{(k,m)},\alpha ^b_{(k',m')} \right] = \omega_{(k,m)}\delta_{k+k'}\delta_{m+m'}\delta^{ab}\nonumber \\
441: \eq
442: and similarly for the $\tilde{\alpha}$'s.
443: The explicit form for the time-dependent part of $X^a$ becomes
444: \beq
445: X^a(\tau,\s,\rho)=x^a+\alpha' p^a \tau + i\sqrt{\frac{\alpha'}{2}}
446: \sum_{(k,m)\neq(0,0)}\frac{e^{iw_{(k,m)}\tau}}{\omega_{(k,m)}}
447: \left[\alpha^a_{(k,m)}e^{ik\s+im\rho}+\tilde{\alpha}^a_{(k,m)}e^{-ik\s-im\rho}
448: \right].
449: \eq
450: The quadratic Hamiltonian (\ref{H0}) takes the form
451: \beq
452: \alpha' H_0= \frac{1}{2}\alpha'(p_9^2+p_{11}^2+p_i^2)+\frac{R_9^2
453: w_0^2}{2\alpha'}+{\cal H},
454: \eq
455: with $i=1,\ldots,8 $ and where ${\cal H}$ contains the contributions from the oscillators,
456: \beq
457: {\cal H} = \frac{1}{2} \sum_{m,k}^\infty \bigl[\alpha^a_{(-k,-m)}\alpha^a_{(k,m)}+
458: \tilde{\alpha}^a_{(-k,-m)}\tilde{\alpha}^a_{(k,m)}\bigr].
459: \eq
460: Letting the momenta in the $X^9,X^{11}$ directions be
461: $p_9=\frac{l_9}{R_9}$ and $p_{11}=\frac{l_{11}}{R_{11}}$, with $l_9,l_{11}$
462: integers,
463: the {\it nine-dimensional} mass operators becomes
464: \beq
465: \label{mass}
466: M^2 =2\, H - p_i^2 =
467: \frac{l_9^2}{R_9^2} + \frac{l_{11}^2}{R_{11}^2} + \frac{R_9^2 w_0^2}{\alpha'^2}
468: + \frac{2}{\alpha'}\,{\cal H}.
469: \eq
470: Clearly, from the eleven-dimensional point of view $\frac{l_9}{R_9}$ and $\frac{l_{11}}{R_{11}}$
471: are just momenta, but they play the role of mass terms in nine
472: dimensions.
473: Schwarz \cite{Schwarz:1995du} showed that the
474: non-oscillating part of the spectrum,
475: $M_0\equiv \frac{l_9^2}{R_9^2} + \frac{l_{11}^2}{R_{11}^2} + \frac{R_9^2 w_0^2}{\alpha'^2}$,
476: matched the corresponding spectrum of the (non-oscillating) type IIB $(p,q)$ string.
477:
478: As in the case of the string, the level matching conditions
479: for the membrane are obtained from the global contraints
480: \beq
481: P^{(\sigma)}=\frac{1}{2\pi\alpha'}\, \int_0^{2\pi} d\sigma \; \p_\sigma X^a \dot{X}^a
482: \equiv 0, \;\;\;\;\;
483: P^{(\rho)}=\frac{1}{2\pi\alpha'}\, \int_0^{2\pi} d\rho \; \p_\rho X^a
484: \dot{X}^a \equiv 0.
485: \eq
486: These can be re-written in terms of mode operators in the
487: following way,
488: \beq
489: N^+_\s -N^-_\s = w_9 l_9, \;\;\;\;N^+_\rho -N^-_\rho = w_{11} l_{11},
490: \eq
491: where
492: \bea
493: N^+_\s &=& \sum_{m=-\infty}^{\infty}\sum_{k=1}^{\infty}\frac{k}{\omega_{km}}
494: \, \alpha^i_{(-k,-m)}\alpha^i_{(k,m)}, \; \; \;
495: N^-_\s =\sum_{m=-\infty}^{\infty}\sum_{k=1}^{\infty}\frac{k}{\omega_{km}}
496: \, \tilde{\alpha}^i_{(-k,-m)}\tilde{\alpha}^i_{(k,m)}, \nonumber \\
497: N^+_\rho &=& \sum_{m=1}^{\infty}\sum_{k=0}^{\infty}
498: \frac{m}{\omega_{km}}\,
499: \bigl[\alpha^a_{(-k,-m)}\alpha^a_{(k,m)}+
500: \tilde{\alpha}^a_{(-k,m)}\tilde{\alpha}^a_{(k,-m)}\bigr] , \; \; \;
501: \nonumber \\
502: N^-_\rho &=& \sum_{m=1}^{\infty}\sum_{k=0}^{\infty}
503: \frac{m}{\omega_{km}}\,
504: \bigl[\alpha^a_{(-k,m)}\alpha^a_{(k,-m)}+
505: \tilde{\alpha}^a_{(-k,-m)}\tilde{\alpha}^a_{(k,m)}\bigr].
506: \ea
507:
508: We are interested in considering states that are BPS.
509: If we want to add a wave to the 2-brane background while preserving
510: supersymmetry, we must align it \cite{Russo:1996if} along the momentum
511: direction. Furthermore, for a BPS state one can have only
512: right-moving oscillations.
513: For the special case $l_{11}=0$, these conditions imply
514: that the oscillations are only along the $\sigma$
515: direction, $N_\rho^\pm=0$, and that there are no left-moving oscillators, $N_\s^-=0$.
516: The condition $N_\rho^+=0$ implies that these states are built by
517: applying $\alpha^i_{(-k,0)}$ to the vacuum, which gives $\omega_{km}\rightarrow w_{11} k$.
518: Thus, one finds
519: \beq
520: {\cal H}=w_{11}\,N_\s^+=w_0 l_9,
521: \eq
522: yielding
523: \beq
524: M^2_{BPS}=\frac{l_9^2}{R_9^2}+ \frac{R_9^2 w_0^2}{\alpha'^2}
525: + \frac{2}{\alpha'}\,{w_0 l_9}=\Bigl(\frac{l_9}{R_9}+4\pi^2 T_3\,w_0\,R_9 \, R_{11}\Bigr)^2.
526: \eq
527: For a general state with $l_9,l_{11}\neq 0$, the constraints become \cite{Russo:1996if}
528: \beq
529: \label{constraints}
530: N_\s^+=l_9 w_9, \;\;\;\;\; N_\rho^+=l_{11} w_{11},
531: \eq
532: giving
533: \beq
534: {\cal H}=w_0 \sqrt{l_9^2+l_{11}^2\tau_2^2}.
535: \eq
536: Thus, we obtain the more general mass formula
537: \beq
538: \label{GeneralSpectrum}
539: M^2_{BPS}=\Bigl(\sqrt{\frac{l_9^2}{R_9^2}+\frac{l_{11}^2}{R_{11}^2}}
540: + 4\pi^2 T_3 w_0 R_9 R_{11}\Bigr)^2,
541: \eq
542: matching the spectrum of the oscillating $(p,q)$ string \cite{Russo:1996if}, as
543: anticipated by Schwarz.
544: One can rewrite the mass spectrum in a more convenient form,
545: \beq
546: \label{massR11mod}
547: M^2_{BPS}=m_P^2 \Bigl(\sqrt{\frac{l_9^2}{R_9^2}+\frac{l_{11}^2}{R_{11}^2}}
548: +w_0 R_9 R_{11}\Bigr)^2,
549: \eq
550: where $m_P$ is the Planck mass, and the radii are now dimensionless.
551: We are particularly interested in the case $l_9=0$, yielding
552: \beq
553: \label{massR11}
554: M^2_{BPS}=m_P^2 \Bigl(\frac{l_{11}}{R_{11}}+w_0 \,R_9
555: \,R_{11}\Bigr)^2.
556: \eq
557: We would like to point out that, since the particle numbers $N_\s^+,
558: N_\rho^+$ must be positive, the constraints (\ref{constraints}) imply that $w_9,
559: l_9$ must have the same sign (and similarly for $w_{11},l_{11}$).
560: Thus, the states described by (\ref{GeneralSpectrum}) are always
561: massive \footnote{We would like to thank K. Hori for discussions regarding this point.}
562: for non-trivial quantum numbers.
563:
564: As we will show in Section \ref{ClTrapping}, such states can
565: play an interesting role on cosmological evolution.
566: However, we are particularly interested in the cosmological consequences of
567: states having a mass of the form
568: \beq
569: \label{ourmass}
570: M^2\sim m_P^2 \Bigl(\frac{|l_{11}|}{R_{11}}-|w_0| R_9 \,R_{11}
571: \Bigr)^2,
572: \eq
573: which can become {\it massless}, and signal an enhancement of symmetry at $R_9\sim R_{11}\sim 1$ (for the
574: case $|l_{11}|=|w_0|)$.
575:
576: Such states do {\it not} appear in the supermembrane spectrum derived in
577: \cite{Russo:1996if}, however one may expect them to be present in the spectrum of M-theory,
578: possibly in heterotic M-theory. In fact,
579: heterotic string theory contains
580: states which become massless at enhanced symmetry points, much in the same way as
581: in the bosonic string case.
582: The mass spectrum of the heterotic string with a compact dimension
583: of radius $R$ is of the form
584: \beq
585: \frac{1}{4}M_{\,h}^{\,2}=\Bigl(\frac{p_L^{\,2}}{2}+N_L-1 \Bigr) + \Bigl(\frac{p_R^{\,2}}{2}+N_R-c_R \Bigr)
586: \eq
587: where $p_{L,R}=n/R \pm w \, R$, and
588: for clarity we have separated the contributions from the
589: left-moving and right-moving sides.
590: For the NS sector $c_R=\frac{1}{2}$, while for the R sector
591: $c_R=0$. Thus, the presence of the zero-point energy allows for
592: states of the form
593: \beq
594: \label{radion}
595: M^{\,2} \sim m_s^2\Bigl(\frac{1}{R}-R\Bigr)^2
596: \eq
597: with $R$ the radius of any one of the six compact dimensions, i.e.
598: the {\it radion}.
599: Thus, the heterotic string shows enhancement of symmetry as $R$ approaches the
600: self-dual radius.
601: At the moment we are not aware of any explicit
602: construction of states of the form (\ref{ourmass}), where the
603: pivotal role is played by the {\it eleven-dimensional radius}, i.e. the
604: dilaton. However, in analogy with what happens in the case of the
605: radion, eq. (\ref{radion}), we expect
606: that states of the form (\ref{ourmass})
607: should be found in M-theory on $S^1/\mathbb{Z}_2$.
608:
609: The lack of such configurations in the derivation of \cite{Russo:1996if}
610: is due to the absence of the zero-point energy in
611: the supermembrane quantization, which is believed to be cancelled
612: by fermionic contributions \cite{Duff:1987cs}.
613: It is also consistent with the fact that such (potentially
614: massless) states are not found in type IIA or IIB string theory, which are
615: obtained from dimensional reduction of 11D M-theory.
616: However, we should mention that supermembrane quantization is still not
617: well understood, and that a full quantum theory of the membrane has
618: proven very difficult to obtain.
619: If, in fact, the zero-point energy were not entirely
620: cancelled by fermions, or if supersymmetry was broken, one
621: could still be able to find states of the form
622: (\ref{ourmass}).
623: Another approach would be to relate
624: known instances of enhanced symmetry to M-theory, through
625: appropriate use of the available dualities, dimensional reduction (and oxidation).
626: Lastly, we mention that considering the supermembrane on backgrounds other than the torus may also offer a resolution,
627: noting for example that Type II strings on K3 (dual to heterotic on $T^6$) also possess enhanced symmetry.
628:
629: Finding whether enhanced symmetry states of the form (\ref{ourmass}) can be present is an interesting problem in its own
630: right, which we would like to consider in the future.
631: However, for the purpose of this paper, we will simply postulate that
632: these states should exist and restrict ourselves to
633: the study of their cosmological consequences.
634:
635: \section{Classical Approach to Trapping \label{ClTrapping}}
636:
637: In the last section we found a set of dyonic solutions
638: whose mass depends explicitly on the dilaton or, from the $11$D
639: perspective, $R_{11}$:
640: \beq
641: \label{heavymass}
642: M^{\,2}=m_p^2\Bigl(\frac{l_{11}}{R_{11}}+w_0 R_9 R_{11}\Bigr)^2.
643: \eq
644: From now on we will set $m_p=1$, and restore explicit $m_p$
645: dependence only when needed.
646: We will now consider the effect of a gas of these dyonic states
647: treated as {\it classical} sources for the vacuum field equations
648: of $11$D SUGRA. The low-energy effective action for the bosonic
649: degrees of freedom is
650: \beq \label{11daction}
651: S=\frac{1}{2\kappa_{11}^2}\int d^{11}x \sqrt{g}
652: \left[R-\frac{1}{48} G_4^2\right]+\frac{1}{6} \int A_3 \wedge G_4
653: \wedge G_4,
654: \eq
655: where $G_4$ is the anti-symmetric four-form flux
656: $G_4=dA_3$, and $\kappa_{11}^2=8\pi G_{11}$, with $G_{11}$ the
657: eleven-dimensional Newton constant. Anticipating a universe with
658: three large spatial dimensions we consider the following ansatz
659: for the metric and four-form
660: \bea \label{ansatz}
661: ds^2&=&-e^{2A}d\eta^2+e^{2B}d\vec{x}\,^2+e^{2C}d\vec{y}\,^2+e^{2D}dz^2,\\
662: G_4&=& h \, dx^1 \wedge dx^2 \wedge dx^3 \wedge dz,
663: \ea
664: where $\vec{x}=(x_1,x_2,x_3), \, \vec{y}=\{y_m\}, \, m=4,\ldots,9$, and
665: $z$ denotes the coordinate of the eleventh dimension. Notice also that
666: we have introduced exponential scale factors $R_m=e^{\,C}$ and
667: $R_{11}=e^{\,D}$. The functions $A,B,C$ and $D$ depend only on
668: $\eta$, and $h$ is a constant. For this choice of flux both the
669: equation of motion and the Bianchi identity for the gauge field
670: are trivially satisfied. The remaining equations of motion follow
671: from varying the action with respect to the full metric $g_{MN}$,
672: with $M,N=(0,1,\ldots,9,11)$, \beq
673: R_{MN}-\frac{1}{2}g_{MN}R=\kappa_{11}^2 \left(
674: \frac{1}{12}F_{MOPQ}F_N^{\; \;
675: OPQ}-\frac{1}{96}g_{MN}F^2+T_{MN}^{\; \text{sources}} \right), \eq
676: where the dyon sources (\ref{heavymass})
677: are included through their stress tensor $T_{MN}^{\; \text{sources}}$.
678:
679: Working in the ideal gas approximation the energy density of these
680: sources can be found from the mass, and is given by
681: \beq
682: \rho=\frac{M}{V}=e^{-3B-6C-D} \left( l_{11} e^{-D}+w_0 e^{C} e^{D}
683: \right).
684: \eq
685: The pressures are then given by
686: \beq \label{pressures}
687: p_{\,3}=-\frac{1}{3}\frac{\partial \rho}{\partial B}-\rho, \;\;\;\;
688: p_{\,6}=-\frac{1}{6}\frac{\partial \rho}{\partial C}-\rho, \;\;\;\;
689: p_{\,11}=-\frac{\partial \rho}{\partial D}-\rho,
690: \eq
691: yielding for the eleven-dimensional pressure
692: \beq
693: p_{\,11}=e^{-3B-6C-D} \left( l_{11} e^{-D}-w_0 e^{C} e^{D}
694: \right).
695: \eq
696: If we take the six dimensions to be at the self-dual radius (i.e. $R_m=1$, or $C=0$)
697: we find that the above pressure will act to stabilize $R_{11}$, corresponding
698: to driving $D \rightarrow 0$.
699: There, the pressure vanishes (for the case $l_{11}=w_0$), and the energy
700: density has a local minimum.
701: We should note that in this analysis the flux would not play a major role, since we are
702: interested in solutions with $B(\eta)$ growing large, and one
703: finds that the flux
704: \footnote{However, the flux will be vital at early times (small $B$) \cite{Friess:2004zk}.}
705: scales as $\sim h^2e^{-6B}$.
706:
707: Such classical attractors have been studied in many contexts throughout the literature.
708: Examples include gases of wrapped strings and branes
709: (see \cite{Battefeld:2005av,Brandenberger:2005nz,Brandenberger:2005fb, Patil:2005nm}
710: and references within), blackhole
711: attractors \cite{Kaloper:2004yj}, D-brane systems \cite{Abel:2005jx}, and
712: cosmologies involving conifold and flop
713: transitions, e.g., \cite{Mohaupt:2005pa}. Despite these elegant
714: ideas for stabilizing moduli, a detailed analysis shows that initial conditions
715: play a crucial role. In the example we have here, one finds that
716: stabilization of $R_{11}$ can be quite generic for fixed $C$.
717: However, as noticed in \cite{Berndsen:2005qq},
718: attempting to stabilize all dimensions simultaneously proves
719: difficult and requires extreme fine-tuning.
720: But perhaps a more serious objection is the {\it validity} of this classical approach.
721: The states (\ref{heavymass}) that we have considered thus far are very heavy for generic values
722: of the moduli. In fact, even at the stabilization point, the energy of these states
723: is typically near the Planck scale.
724: Thus, the low-energy effective theory that provided us with the equations of motion is no
725: longer the proper
726: framework, given the inclusion of massive states. Furthermore, one
727: cannot arbitrarily add only some, but not all, of the massive
728: states of the strings or branes.
729:
730: We will turn to a more self-consistent approach in the next section, but it is important
731: to stress that, although the classical approach of this section was naive,
732: it is still interesting to see how the inclusion of truly {\em stringy} states
733: can affect the dynamics.
734: That is, low-energy SUGRA is {\it not} string theory, and the addition of the
735: membrane winding and oscillator contributions can be considered a first correction
736: toward the UV completion of the theory. Thus, a better
737: understanding of the full string (or membrane) spectrum is a first
738: step towards addressing the string vacuum problem and exploring
739: string phenomenology.
740: However, we do support the viewpoint that this approach is not
741: truly consistent, and whereas in the next section we will find a self-consistent way to
742: include {\em stringy} states in the low-energy theory.
743:
744: \section{Membrane Production in Time-Dependent Backgrounds \label{QmTrapping}}
745:
746: In this section we consider the corrections which may result from the dynamics
747: of moduli in the low-energy effective action of string (M-)theory, namely SUGRA.
748: As we have seen, string models contain states whose masses depend explicitly on
749: massless moduli.
750: If the moduli are time-dependent, the masses of such string
751: states change, and can even vanish.
752: Of course, this mechanism is not specific to string theory;
753: the Higgs is a field theory example of a modulus which controls
754: particle masses, however the crucial difference here is the target space of the string is space-time itself.
755:
756: Typically the states become light as the moduli approach
757: locations where symmetries play a special role (ESPs).
758: Near such points, the states can be quantum produced, and must
759: therefore be incorporated into the effective field
760: theory description.
761: As an example of this {\em stringy} Higgs mechanism, consider the case of a D-brane wrapping a collapsing cycle of the extra dimensions.
762: It was realized some time ago \cite{Strominger:1995cz}, that by including additional light states resulting from the
763: collapse of a wrapped D-brane on a shrinking cycle of the Calabi-Yau, the naive singularities of the conifold geometry could be resolved.
764: The singularity in the moduli space was realized as an artifact of integrating out degrees of freedom that were no longer massive and beyond the cutoff
765: of the effective theory. We see that dynamics can play a vital role when consider which degrees of freedom to include in the effective theory,
766: since the mass scale can be dynamical given the evolution of background fields (in this case the collapse of the extra dimensions and the brane).
767:
768: One can identify several corrections to the moduli space approximation
769: resulting from the inclusion of the new light degrees of freedom.
770: Although, in this paper we will concentrate on on-shell production
771: neglecting other types of corrections, which we will mention
772: briefly in the conclusions.
773: As we will see, particle production will be suppressed generically,
774: but it will be non-zero and significant for the enhanced symmetry
775: states mentioned at the end of Section \ref{MthStates}.
776: We will then study the effects of the resulting backreaction on
777: cosmological evolution, and show that it can lead to trapping of
778: the moduli under consideration.
779: Such a trapping mechanism has been previously studied in \cite{Watson:2004aq,Kofman:2004yc}.
780:
781: \subsection{Time-Dependent Backgrounds}
782:
783: Let us begin by considering the low energy effective theory for M-theory, i.e. $11$D SUGRA.
784: The effective theory for the massless bosonic fields is given by the action (\ref{11daction}).
785: A class of solutions was found in \cite{Lidsey:1999mc}
786: \bea
787: \label{FullScaleFactors}
788: ds^2&=&-e^{2A}d\eta^2+e^{2B}d\vec{x}\,^2+e^{2C}d\vec{y}\,^2+e^{2D}dz^2, \nonumber \\
789: G_4&=& h \, dx^1 \wedge dx^2 \wedge dx^3 \wedge dz,
790: \eea
791: with the scale factors given by
792: \bea
793: A&=&3B+D \nonumber \\
794: B&=&B_0+\frac{1}{3}\log{\sec{h\eta}}+
795: \frac{q}{2}\log{\frac{1+\sin{h\eta}}{\cos{h\eta}}} \nonumber \\
796: C&=&C_0+\frac{1}{6}\log{\cos{h\eta}} \nonumber \\
797: D&=&D_0-\frac{1}{3}\log{\cos{h\eta}}+\frac{1-3q^2}{6q}
798: \log{\frac{1+\sin{h\eta}}{\cos{h\eta}}},
799: \ea
800: where $B_0,C_0,D_0 \; \text{and} \; |q| \leq 1/\sqrt{3}$ are real constants and $-\frac{\pi}{2h} \leq \eta \leq \frac{\pi}{2h}$.
801: Notice that in the limit of no flux, $h=0$, we recover flat space and $-\infty \leq \eta \leq \infty$.
802: We also note that, although this space-time contains a singularity (located at
803: $-\frac{\pi}{2h}$ for $q>0$),
804: we will only be interested in the evolution away from the singularity
805: \footnote{In \cite{Friess:2004zk}, winding mode creation in this background was studied
806: in order to try to resolve the cosmological singularity.
807: In this paper, however, we will focus on production away from the singularity.}.
808:
809: We now want to consider the production of membrane states
810: of the type discussed in Section \ref{MthStates},
811: having an effective mass
812: which depends on the radii of the compact dimensions.
813: The first question one may ask is why we expect any production at all, since
814: at low energy scales production of string and membrane states should be strongly
815: suppressed.
816: For the moment let us push forward, keeping this important question in mind.
817: Consider the quantum equation of motion for a string/membrane state labelled by $\chi$.
818: As discussed in \cite{Lawrence:1995ct,Gubser:2003vk} the string constraint equations amount to a
819: wave equation for $\chi$,
820: \be \label{chieom}
821: {\chi}^{\prime \prime}_k+\omega_k^2 \chi_k=0,
822: \ee
823: where a prime denotes the derivative with respect to $\eta$, and $\chi_k$
824: are the Fourier modes of $\chi$.
825: We have removed the friction term by a field redefinition $\chi \rightarrow \chi e^{-3C}$.
826: For simplicity, we will rescale time $\eta \rightarrow \eta h^{-1}$ so that
827: $-\frac{\pi}{2}< \eta < \frac{\pi}{2}$.
828: The time dependent frequency is then given by
829: \be \label{omega}
830: \omega^2=e^{6B+2D}h^{-2}\left( \frac{k_3^2}{e^{2B}} + \frac{k_6^2}{e^{2C}} + \frac{k_{11}^2}{e^{2D}} +m^2(\eta) +\xi R -e^{-6B-2D}h^2(9C^{\prime \, 2}
831: +3C^{\prime \prime}) \right),
832: \ee
833: where $R$ is the $11$D Ricci scalar, $\xi$ is the coupling of $\chi$ to the
834: space-time background, and $m^2$ represents the contributions to the mass coming
835: from the winding and oscillations of the membrane.
836: Also note that $k_3, k_6$ and $k_{11}$ denote {\it comoving}
837: momentum.
838: It will be more convenient at times to think in terms of the nine-dimensional mass
839: \bea \label{masseff}
840: m^2_{{eff}}&=&\left( k^2_{11}e^{-2D} +m_w^2+m_{\text{osc}}^2 \right) \nonumber \\
841: &=&m_p^2 \left( l_{11}e^{-D}-\omega_0 e^D e^C \right)^2,
842: \eea
843: since it is when this mass vanishes that we expect significant production.
844: Choosing $B_0=C_0=D_0=0$ for simplicity, these states
845: will become massless (for the case $l_{11}=w_0$) at $\eta=0$, where $D=C=0$.
846: Given the time-dependent frequency (\ref{omega}), with $l_{11}$ and $w_0$
847: left arbitrary, we are now ready to consider the production of the
848: membrane states.
849:
850: \subsection{The Steepest Decent Method \label{sdm}}
851:
852: Before considering the case of interest, let us review the
853: standard method for calculating particle production of states with
854: a time-dependent frequency (see e.g. \cite{Birrell:1982ix}). A
855: formal solution to the mode equation (\ref{chieom}) is given by
856: \beq \chi_k = \frac{\alpha_k}{\sqrt{2\omega_k}}e^{-i\int \omega_k
857: d\eta}+\frac{\beta_k}{\sqrt{2\omega_k}}e^{i\int \omega_k d\eta}.
858: \eq The normalization condition for the scalar field is then
859: $|\alpha_k(\eta)|^2-|\beta_k(\eta)|^2=1$, which can be used to
860: write the equation of motion as \bea \alpha_k' &=&
861: \frac{\omega_k'}{2\omega_k} e^{2i\int
862: \omega_k(\eta')d\eta'}\beta_k \nonumber \\ \beta_k' &=&
863: \frac{\omega_k'}{2\omega_k} e^{-2i\int
864: \omega_k(\eta')d\eta'}\alpha_k. \ea Initially we start near the
865: adiabatic vacuum and we have $\beta_k<<1$ and $\alpha_k \sim 1$ so
866: that \beq \label{beta} \beta_k \sim \int d\eta
867: \frac{\omega_k'}{2\omega_k} e^{-2i\int^\eta
868: \omega_k(\eta')d\eta'}, \eq to leading order. Since $|\beta|^2$
869: gives the number of particles produced, we see that production is
870: suppressed as long as $\omega^\prime \ll \omega$. Conventionally,
871: one defines the dimensionless non-adiabatic parameter
872: ${\omega^\prime}/{\omega^2}$, indicating that particle production
873: becomes significant when
874: \be
875: \frac{\omega^\prime}{\omega^2} \sim 1.
876: \ee
877: In order to calculate the total amount of particles produced we need to estimate the integral in
878: (\ref{beta}). A method for approximating this integral was found
879: in \cite{Chung:1998bt}, which we will summarize below.
880:
881: We will be interested in $\omega_k$ of the form $\omega_k=\sqrt{k^2+m^2 f(\eta)}$,
882: where we absorb any time dependence of the background into $f(\eta)$.
883: The poles of the integrand in (\ref{beta}) are also
884: branch points in the complex $\eta$ plane. Let us
885: assume for simplicity that $\omega_k$ has a single pole at
886: $\eta^\star$.
887: Then, near the branch point, the integral in the exponent of (\ref{beta})
888: can be expanded in the following way:
889: \beq
890: \int_{\eta_{in}}^\eta \omega_k(\eta') d\eta' =\int_{\eta_{in}}^\eta \omega_k(\eta')
891: d\eta' + \frac{2M}{3}\sqrt{f'(\eta^\star)}(\eta-\eta^\star)^{3/2},
892: \eq
893: keeping the leading term in the expansion of $f(\eta)$ about $\eta^\star$.
894: We then find
895: \beq
896: \beta_k = \Bigl( \frac{1}{4}\int_{f^\star}\frac{d\delta}{\delta}e^{\frac{-4iM}{3}
897: \sqrt{f'(\eta^\star)}\delta^{3/2}} \Bigr) \,
898: e^{-2i\int_{\eta_{in}}^{\eta^\star} \omega_k(\eta')d\eta'},
899: \eq
900: where the $f^\star$ denotes the steepest descent contour.
901: The factor in front of the exponential can be shown to equal
902: $\frac{i\pi}{3}$, giving us
903: \beq
904: \beta_k = \frac{i\pi}{3} e^{-2i\int_{\eta_{in}}^{r} \omega_k(\eta')d\eta'} \,
905: e^{-2i\int_{r}^{\eta^\star} \omega_k(\eta')d\eta'},
906: \eq
907: where $\eta^\star=r-i\mu$.
908: The first integral in the expression
909: above is real, making the first exponential a pure phase.
910: The second integral can be approximated by
911: $\int_{r}^{\eta^\star} \omega_k(\eta')d\eta' \sim i \pi \mu
912: \omega_k(r)$, yielding
913: \beq \label{betaform}
914: |\beta_k|^2 = \biggl(\frac{\pi}{3}\biggr)^2 e^{-\pi\mu\omega_k(r)}.
915: \eq
916:
917: Thus, to calculate the leading contribution to particle production
918: it is sufficient to identify the real and imaginary contributions
919: to the zeros of $\omega_k$. Because of the choice of integration
920: contour, we should note that we are only interested in the zeros which are in the
921: lower half plane.
922:
923: \subsection{Membrane Production}
924:
925: We now want to determine which conditions will lead to significant production
926: in the chosen background (\ref{FullScaleFactors}). The
927: adiabatic approximation is valid as long as ${\omega^\prime}/{\omega^{2}}<1$,
928: and as long as this condition holds production will be insignificant.
929: Thus, the first step is to examine when the non-adiabaticity is appreciable for our frequency
930: \be \label{ourfreq}
931: \omega^2_k=e^{6B+2D}h^{-2}\left( \frac{k_3^2}{e^{2B}} +m_{eff}^2(\eta)
932: +\xi R -e^{-6B-2D}h^2(9C^{\prime \, 2}+3C^{\prime \prime}) \right),
933: \ee
934: where we will neglect any momentum in the extra dimensions (i.e. we take $k_6=0$),
935: but keep $k_{11}$ non-zero.
936:
937: One obvious location where non-adiabaticity becomes large is near the cosmological
938: singularity ($\eta=-\pi/2$), which was the case examined in \cite{Friess:2004zk}.
939: Here, however, we are not interested in the region near the
940: singularity. Thus, as long as we are not too close to the singular
941: region, we can still construct the appropriate in-vacuum
942: \footnote{This is analogous to the calculation of production at the end of inflation,
943: where the initial big-bang singularity is irrelevant to the calculation.
944: See \cite{Birrell:1982ix} for a detailed discussion on the method of constructing
945: an asymptotic, adiabatic vacuum in such singular space-times.}.
946: Let us now examine the behavior of the non-adiabatic parameter away from
947: the initial singularity. We expect that away from the values $C=D=0$ the mass term
948: should dominate over all other terms in the frequency (\ref{ourfreq}).
949: In this limit, $$\frac{\omega^\prime}{\omega^2} \approx \frac{m_{eff}^\prime}{m_{eff}^2},$$
950: and we expect the non-adiabaticity to be peaked near $m_{eff}\approx 0$, which is where
951: the states become massless.
952:
953: In Figures \ref{fig1}, \ref{fig2}, and \ref{fig3}, we present the behavior of the
954: {\it exact} non-adiabatic parameter for various values of the background and momenta.
955: In all cases we find that the production is sharply peaked around $\eta \approx 0$.
956: As noted in \cite{Kofman:2004yc} (and references within), this means that production can
957: be treated as an instantaneous event, making the calculation of production
958: and its backreaction much more tractable. We see from the figures that
959: away from $\eta \approx 0$ the adiabatic vacuum is an excellent approximation,
960: and that well-defined in and out regions do indeed exist.
961:
962:
963: \begin{figure}
964: \includegraphics[width=10.5cm]{fig1}
965: \caption{\label{fig1} The non-abiabatic parameter ${\omega^\prime}/{\omega^{2}}$
966: for different values of $\xi$, with
967: $\xi=0$ (top curve), conformal coupling $\xi=9/40$ (middle curve), and $\xi=5$ (bottom curve).
968: As $\xi$ increases, non-adiabaticity is suppressed.
969: Note that the non-adiabticity is
970: concentrated near $\eta \approx 0$ for all choices of $\xi$.}
971: \end{figure}
972: %
973: \begin{figure}[!]
974: \includegraphics[width=9.5cm]{fig2}
975: \caption{\label{fig2} The non-adiabatic parameter ${\omega^\prime}/{\omega^{2}}$ for different
976: values of the flux $h=1/150$ (top curve), $h=1/100$ (middle curve), and $h=1/50$ (bottom curve).}
977: \end{figure}
978: %
979: %
980: \begin{figure}[!]
981: \includegraphics[width=9.5cm]{fig3}
982: \caption{\label{fig3} The non-adiabatic parameter ${\omega^\prime}/{\omega^{2}}$ for different values of the
983: momentum $k_3=1/500$ (top line), $k_3=1/100$ (middle line), and $k_3=1/50$ (bottom line).}
984: \end{figure}
985: %
986:
987: One point of concern is the effect of the time-dependent geometry on the
988: effective mass $$m^2_{total}=m_{eff}^2+\xi R + \text{gravity terms}$$ through the
989: last three terms in (\ref{ourfreq}).
990: As seen in Fig. \ref{fig1}, the terms including the coupling to the geometry $\xi$,
991: have the effect of suppressing the amount of production, but this suppression is
992: negligible for reasonable values of the coupling. In particular, the cases of minimal
993: ($\xi=0$) and conformal $(\xi=9/40)$ coupling have little or no effect. This can be understood
994: by examining the Ricci curvature which, for small flux, is negligible away from the
995: singular region (i.e., $R \sim h^2$). The requirement of small flux can be seen
996: from Fig. 2, where we find that production will be suppressed unless $h<m_p$.
997: Fig. 3 shows a similar result for the momenta, $k_3 < m_p$.
998: Let us now consider the production from a more quantitative perspective.
999:
1000: We have seen that the non-adiabaticity is focused near $\eta \approx 0$. Using this
1001: and the effective mass (\ref{masseff}), we can expand the frequency (\ref{ourfreq})
1002: for small $\eta$
1003: \be \label{approxomega}
1004: \omega_k^2(\eta)=f_0+f_1 \eta+f_2 \eta^2,
1005: \ee
1006: where, after restoring the explicit dependence on $m_p$ ,
1007: \bea
1008: f_0&=& \frac{k_3^2+a_0m_p^2+\frac{1}{2} h^2}{h^2}, \;\;\;\;\;\;\;
1009: f_1 = \frac{a_1 k_3^2+a_2 m_p^2}{h^2}, \;\;\;\;\;\;\;
1010: f_2 = \frac{a_3k_3^2+a_4m_p^2+\frac{1}{4} h^2}{h^2}, \nonumber \\
1011: a_0&=&(w_0-l_{11})^2,
1012: \;\;\;\;a_1=\frac{1+3q^2}{3q}, \;\;\;\;
1013: a_2=\frac{2w_0(w_0-l_{11})+3q^2(w_0^2-4w_0l_{11}+3l_{11}^2)}{3q} \nonumber \\
1014: a_3&=&\frac{1+24q^2+9q^4}{18q^2}, \;\;
1015: a_4=\frac{2w(w-l)+3q^2(13w^2-23wl+6l^2)+q^4(9w^2-72wl+91l^2)
1016: }{18 q^2}.\nonumber
1017: \ea
1018: Following the method outlined in Section \ref{sdm}, we proceed by finding the zeros of (\ref{approxomega}).
1019: The frequency vanishes at
1020: \beq
1021: \eta^*=\frac{-f_1\pm\sqrt{f_1^{\, 2}-4\, f_0 \, f_2}}{2f_2}.
1022: \eq
1023: Remembering that we are interested in the zeros in the lower
1024: half plane, we choose
1025: \beq
1026: \eta^\star=-\frac{f_1}{2f_2}-i\frac{\sqrt{4f_0
1027: f_2-f_1^2}}{2f_2}\equiv r-i\mu,
1028: \eq
1029: where we have $4f_0 f_2-f_1^2>0$.
1030: The amount of particle production for a given mode is then given by (\ref{betaform}),
1031: and is found to be
1032: \beq
1033: |\beta_k|^2 = \biggr(\frac{\pi}{3}\biggr)^2 e^{-\pi\frac{4f_0
1034: f_2-f_1^2}{4 \, f_2^{3/2}}}.
1035: \eq
1036: Moreover, using the fact that $m_p\gg k_3,h$, we find
1037: \beq \label{betaeqn}
1038: |\beta_k|^2 = \biggr(\frac{\pi}{3}\biggr)^2 e^{-\frac{A_1\,m_p}{h}}
1039: e^{-\frac{A_2\,k_3^2+A_3\,h^2}{m_p h}+{\cal
1040: O}(\frac{1}{m_p^2})},
1041: \eq
1042: where $A_1, A_2, A_3$ are constants that depend on $w_0, l_{11},
1043: q$.
1044:
1045: Given this expression we can now return to the initial concern that string and
1046: membrane particle production should suffer Planck scale suppression
1047: at low energies. This can be seen by the suppression factor $e^{-\frac{A_1\,m_p}{h}}$
1048: appearing in (\ref{betaeqn}), which is related to the usual suppression
1049: due to the Hagedorn density of states (see e.g. \cite{Gubser:2003vk}).
1050: However, in this case we find that $A_1 \propto (l_{11}-w_0)^3$, which vanishes
1051: when $l_{11}=w_0$, consistent with the condition for having
1052: massless states\footnote{This is the motivation for wanting states
1053: having a mass of the form (\ref{ourmass}).}
1054: at the self-dual radius $D=0$. Thus, the enhanced symmetry results in
1055: additional light states which will be copiously produced.
1056: In fact, we find significant particle production when $l_{11}=w_0$,
1057: given by
1058: \beq \label{betaeqna}
1059: |\beta_k|^2 = \biggr(\frac{\pi}{3}\biggr)^2 e^{-\frac{3\pi}{2\,m_p\,h}
1060: \bigl(\frac{h^2+2 k_3^2}{l_{11}\gamma}
1061: \bigr)},
1062: \eq
1063: where $\gamma\equiv \frac{1-3q^2}{q}$.
1064: Furthermore, we will choose $l_{11}=w_0=1$, since states
1065: with higher winding and momentum numbers would decay to this configuration.
1066: We can immediately see from (\ref{betaeqna}) the same behavior that we found for the non-adiabatic parameter in the previous section.
1067: Namely, we see that for $h, k_3 < m_p$ production will result and above the Planck scale production will cease. This behavior agrees with the numerical studies
1068: of the non-adiabatic parameter, which can be seen in Fig. \ref{fig2} and Fig. \ref{fig3}.
1069:
1070: From (\ref{betaeqna}) we can calculate the eleven-dimensional number
1071: density at the creation time $\eta^*$ by summing over all modes
1072: \beq \label{densities}
1073: n_\chi (\eta^*)
1074: =\frac{1}{V_0^{(7)}} \int_0^\infty \frac{d^3k_3}{(2\pi)^3}
1075: \; |\beta_k|^2,
1076: \eq
1077: $V_0^{(7)}$ is the dimensionful volume of the extra dimensions
1078: at the time the particles were produced \footnote{We
1079: note that in the calculation of the 11D densities there is an
1080: implicit integration over the momenta of the extra dimensions (and the Kaluza-Klein modes).
1081: However, we have chosen to work with the $l_{11}=1$ states
1082: and any additional momentum (e.g. $k_6$) is taken to vanish.}.
1083: The energy density of created particles is given by
1084: \beq \label{edense}
1085: \rho= \frac{1}{V_0^{(7)}} \int_0^\infty \frac{d^3k_3}{(2\pi)^3} \; \omega_k
1086: |\beta_k|^2.
1087: \eq
1088: Finally, the pressure can be found from the energy density as in (\ref{pressures}),
1089: and the effective stress energy tensor of the produced membrane states is then
1090: \be \label{st}
1091: T^{M}_{\; \; N}=diag \left( -\rho, p_3, p_6, p_{11} \right).
1092: \ee
1093:
1094: \section{Cosmology, Backreaction, and Moduli trapping \label{cosmo}}
1095:
1096: We now want to consider the effect of the states produced in the last section
1097: on the evolution
1098: of the background (\ref{FullScaleFactors}).
1099: One possible way to include the effects of such states would be
1100: to study the evolution of
1101: linear perturbations about the background (\ref{FullScaleFactors}).
1102: However, one finds that this approach is not adequate, since the presence of the states
1103: alters the evolution drastically, making the effect of order one.
1104: Instead, we consider the new metric
1105: \be \label{newmetric}
1106: ds^2=-dt^2+a(t)^2 dx_3^2 + b(t)^2 dx_6^2+R_{11}^2(t) dx_{11}^2,
1107: \ee
1108: where we will work in synchronous gauge, and we have assumed that the perturbed metric will
1109: respect the symmetries (topology) of the original metric\footnote{
1110: Note that for simplicity we will take six of the radii to be equal (i.e., $R_m=b$
1111: with $m=4,5,6,7,8,9$).
1112: This could easily be generalized without changing our conclusions.}.
1113: Introducing the Hubble parameters $H_3=\frac{\dot{a}}{a}, H_6=\frac{\dot{b}}{b}$ and
1114: $H_{11}=\frac{\dot{R}_{11}}{R_{11}}$, the equations of motion can be written as
1115: \be \label{theeom}
1116: \dot{H}_i=H_i \sum_{j} H_j + 8\pi G_{11} \left( p_i +\frac{1}{10}
1117: \left( \rho- \sum_{j} p_j \right) \right),
1118: \ee
1119: with $i,j=1 \ldots 9,11$ running over all spatial dimensions. We have again taken
1120: the matter sources
1121: to be in the form of a perfect fluid characterized by their energy density and pressures,
1122: and having a stress tensor given by
1123: \be \label{sth}
1124: T_{MN}=diag \left( \rho, a^2 p_3, b^2 p_6, R_{11}^2 p_{11} \right).
1125: \ee
1126:
1127: The constraint equation
1128: \be
1129: \sum_{i<j} H_i H_j =8\pi G_{11} \rho
1130: \ee
1131: and the equations motion allow one to obtain the continuity equation
1132: \be
1133: \dot{\rho}=-\sum_i \left( \rho + p_i \right) H_i.
1134: \ee
1135: In general these equations are very difficult to solve.
1136: However, in the absence of matter (i.e., $\rho=p_i=0$) one finds the well-known
1137: Kasner solutions,
1138: \bea
1139: a(t)\sim t^{c_1}, \;\;\;\;\;\; b(t)\sim t^{c_2}, \;\;\;\;\;\;\; R_{11}(t)\sim t^{c_3},
1140: \nonumber \\
1141: \text{with} \;\;\;\sum_i c_i = \sum_i (c_i)^2=1.
1142: \eea
1143: Thus, in a universe initially filled with radiation or matter the expansion will
1144: dilute the sources,
1145: and we expect the late-time behavior to approach that of the Kasner solution.
1146:
1147: We want to include the effect of the produced states on the background
1148: (\ref{FullScaleFactors}).
1149: As we found in the last section we can include the states through their energy
1150: density (\ref{edense}).
1151: It is important to note that number density is only well defined in the adiabatic
1152: out region.
1153: That is, near the region of non-adiabaticity, particle number (and thus energy density)
1154: can fluctuate.
1155: However, following \cite{Kofman:2004yc} we will assume that after
1156: the first pass through the production point no further production occurs, and the particle
1157: number remains fixed.
1158: This approximation is certainly acceptable, since the exact treatment would merely increase
1159: the number of particles, enhancing the trapping mechanism we will discuss.
1160:
1161: Perhaps a more serious drawback is the validity of the SUGRA approximation near the
1162: production point.
1163: For the background we have considered, production will occur near the Planck scale at the
1164: center of the M-theory moduli space.
1165: From the ten-dimensional point of view, this will be a fixed point of S-duality where a
1166: perturbative description of the theory is not available at this time.
1167: However, we can trust our equations away from the non-adiabatic region, and we are primarily
1168: interested in the attractive behavior towards this region.
1169: Thus, we will consider the effects coming from production, remembering
1170: that further corrections may become important near the non-adiabatic region.
1171: In the worst case, we will see that the corrections we consider lead us naturally to the
1172: non-adiabatic region, where a further knowledge of M-theory dynamics is needed.
1173:
1174: Using the expression (\ref{betaeqna}) in (\ref{densities}) we find for the number density
1175: at the time of creation $\eta^{*}$
1176: \bea
1177: n_\chi(\eta^{*}) &=& \frac{1}{\vol}
1178: \int_0^\infty \frac{d^3k}{(2\pi)^3} \; \left( \frac{\pi}{3} \right)^2
1179: e^{ -\frac{3\pi}{h\gamma m_p} \left( k_3^2+ \frac{1}{2}h^2 \right) } \\
1180: &=& \frac{\sqrt{3}\gamma^{3/2}}{648 \, \pi} \left(m_p \, h\right)^{5}
1181: e^{-\frac{3\pi h}{2 \gamma m_p}},
1182: \eea
1183: where we used $V_0^{(7)}=(h\,m_p)^{-7/2}$.
1184: This is the number density of particles resulting from production
1185: at the time when the scale factors pass near the region of non-adiabaticity $D=C=0$.
1186: %
1187: In terms of the new metric (\ref{newmetric}) the creation time
1188: will be denoted by $t_0$, and the production point corresponds to $R_{11}=b=1$.
1189: Away from the non-adiabatic region the number of particles will remain
1190: constant, and the density in the new metric is given by
1191: \beq
1192: n_\chi(t)=\left( \frac{a^3(t_0) \, b^6(t_0) \, R_{11}(t_0)}{a^3(t)\, b^6(t)\, R_{11}(t)}
1193: \right)n_\chi(t_0)\equiv \frac{V(t_0)}{V(t)}n_\chi(t_0).
1194: \eq
1195:
1196: At this point it is convenient to return to setting
1197: $m_p=1$; we will reinstate the explicit dependence on the Planck
1198: mass when necessary.
1199: The energy density at time $t_0$ is given by
1200: \beq
1201: \rho(t_0) =\frac{\pi^2}{9 \vol} \int_0^\infty \frac{d^3k}{(2\pi)^3} \;
1202: \omega_k \, e^{ -\frac{3\pi}{h \gamma} \left( k_3^2+ \frac{1}{2}h^2 \right) }.
1203: \eq
1204: The frequency $\omega_k$ evaluated in the new background takes the
1205: form
1206: \be
1207: \omega_k^2 \sim \frac{k_3^2}{a^2}+ \left( \frac{1}{R_{11}}-R_{11} b \right)^2
1208: \equiv \frac{k_3^2}{a^2}+ m_{eff}^2,
1209: \ee
1210: where we have dropped gravitational terms which have been shown to be
1211: of higher adiabatic order and are negligible (recall that these terms scale like $h^2$, and
1212: $h^2 \ll m_{eff}^2$).
1213: Using this frequency we find that the energy density at $t_0$ is
1214: given by
1215: \beq
1216: \rho(t_0)= \frac{\pi^2}{9\vol} \int_0^\infty \frac{d^3k}{(2\pi)^3}
1217: \sqrt{\frac{k_3^2}{a^2}+ m_{eff}^2} \;
1218: e^{ -\frac{3\pi}{h \gamma} \left( k_3^2+ \frac{1}{2}h^2 \right)}.
1219: \eq
1220: At a later time $t$, the energy density scales in the same way as
1221: the number density,
1222: \beq
1223: \label{rhogo}
1224: \rho(t)=\frac{V(t_0)}{V(t)} \frac{a \, \gamma \,h^{9/2} }
1225: {144 \pi} \; m_{eff}^2 \; e^{-\frac{3\pi}{2\gamma}\left( h-a^2 h^{-1}
1226: m_{eff}^2\right)} \mathcal{K}_1\left( a^2 h^{-1}m_{eff}^2 \right),
1227: \eq
1228: where $\mathcal{K}_1$ is a Bessel function of the second kind.
1229:
1230: \subsection{Cosmological Evolution and Trapping}
1231:
1232: We are interested in the dynamics after passing through the ESP at $R_{11}=b=1$.
1233: Notice that in the absence of matter the background solution (\ref{FullScaleFactors})
1234: predicts that the radii (moduli) will continue to evolve to larger values.
1235: We will demonstrate that, by including the states produced at the ESP via their stress energy
1236: tensor (\ref{st}),
1237: the motion of the radii will be reversed back towards the ESP, and the moduli can eventually
1238: become trapped.
1239: We again note that there should also be a contribution to the stress tensor coming from the flux,
1240: but such terms scale as $\rho \sim p_i \sim \frac{h^2}{a^6}$ and are therefore only important at
1241: early times (i.e. they are red-shifted by the expansion of $a(t)$).
1242:
1243: The evolution begins in the adiabatic region, where $R_{11}>1$, $b>1$ and the mass term
1244: dominates the energy.
1245: In this limit the energy density (\ref{rhogo}) becomes
1246: \be \label{rhoapprox}
1247: \rho \approx m_{eff} \, n_\chi(t).
1248: \ee
1249: In the opposite limit, near $R,b \approx 1$, the mass is negligible and the energy density becomes
1250: \be \label{rhoapprox2}
1251: \rho \approx \frac{n_\chi(t)}{a}.
1252: \ee
1253: From (\ref{rhoapprox}) we can determine the pressure in the respective dimensions using (\ref{pressures}).
1254: We find that for $R_{11}>1$ and $b>1$ these scale as
1255: \bea
1256: p_{\,3}&\approx&0, \;\;\;\;\;
1257: p_{\,6} \approx-\frac{n_\chi(t)}{6} R_{11}b, \;\;\;\;\;\;
1258: p_{\,11}\approx-n_\chi(t)R_{11}b \;\;\;\;\;\;\;\;
1259: \left( R_{11}>1 \; \text{and} \; b > 1\right).
1260: \eea
1261: We see that for large $R_{11}$ (or large $b$), the negative pressure due to the wrapped M2 branes
1262: dominates.
1263: From the equations of motion (\ref{theeom})
1264: we see that as $R_{11}$ grows large the negative source terms will dominate over
1265: the first terms
1266: on the right side of (\ref{theeom}), given that $H_{11}$ is not too large.
1267: At the turning point $H_{11}=0$ and we find that $\dot{H}_{11}<0$, due to the large negative pressure $p_{11}$,
1268: showing that $R_{11}$ reaches a
1269: maximum value and
1270: turns back toward the ESP \footnote{It is important to note that in anisotropic space-times
1271: negative pressures lead to contraction, not accelerated expansion.}.
1272: After the radii pass through the ESP for a second time, we have $R_{11}<1$, and the
1273: pressures are then given by
1274: \bea \label{pb}
1275: p_{\,3}&\approx&0, \;\;\;\;\;\;\;
1276: p_{\,6}\approx \frac{n_\chi(t)}{6} R_{11}b, \;\;\;\;\;\;\;
1277: p_{\,11}\approx \frac{n_\chi(t)}{R_{11}} \;\;\;\;\;\;\;\;\;\;\;\;\;\;\; \left(R_{11}<1\right),
1278: \eea
1279: In the last step we have assumed that the radii have moved sufficiently below the ESP so that we are again
1280: in an adiabatic region, and (\ref{rhoapprox}) still
1281: gives the relevant energy density\footnote{Strictly speaking this is not quite correct.
1282: As the radii pass back through the ESP (non-adiabatic region), further particle production
1283: is possible and the density of particles can increase. However, including the production of
1284: additional states will only act to enhance the trapping mechanism that we will discuss.
1285: See \cite{Kofman:2004yc} for a related discussion.}.
1286: We see from (\ref{pb}) that as $R_{11}$ continues to evolve towards smaller values, the
1287: pressure becomes large and positive. From
1288: (\ref{theeom}) this means that $R_{11}$ will reach a minimum value and
1289: go back towards the ESP.
1290: This behavior will continue and $R_{11}$ will oscillate around the ESP, with the oscillations
1291: damping due to the expansion of $a(t)$.
1292:
1293: However, one place we have been cavalier is in the evolution of the extra dimensional scale factor $b$.
1294: In fact, from its associated
1295: pressure (\ref{pb}) we see that for $b \ll 1$ the pressure will vanish.
1296: Thus, since there is no pressure to
1297: prevent its collapse, $b$ will continue to run to smaller values.
1298: Furthermore, we have over-simplified the entire evolution by assuming that $R_{11}=1$ and $b=1$
1299: occur simultaneously. If this were the case it would be the result of extreme fine-tuning.
1300: \begin{figure}
1301: \includegraphics[width=10.5cm]{fig4}
1302: \caption{\label{fig4} Evolution of $R_{11}$ and $b$ for the exact density (\ref{rhogo}) and pressures, including only the membrane sources related to $R_{11}$.}
1303: \end{figure}
1304: Instead, as can be seen in Fig. \ref{fig4}, we find that for differing values of $R_{11}$ and
1305: $b$ trapping of $R_{11}$ can occur away from the ESP at a value that is determined by the
1306: running of $b$ to its asymptotic value\footnote{We note that upon compactification to
1307: ten dimensions this is the result found in \cite{Watson:2003gf}, where it was shown that the
1308: radii could be stabilized at the expense of a running dilaton}.
1309:
1310: Despite the evolution of $b$, we can remedy the situation by simply including the other membrane
1311: states discussed in Section \ref{MthStates}. That is, near $b=1/\sqrt{R_{11}}$ there are additional
1312: massless states that can be produced with masses
1313: \be
1314: m^{(b)}_{eff}= \left| \frac{1}{b}-R_{11}(t) b(t) \right|,
1315: \ee
1316: where now the momentum is taken in the $b$ direction (i.e., $k_9 \neq 0$, $k_{11}=0$).
1317: The production of these states is handled analogously
1318: to the previous states, and leads to an additional contribution to the energy density
1319: \be
1320: \rho^{(b)}=6 \, \tilde{n}_\chi(t) \, m_{eff},
1321: \ee
1322: where the factor of six comes from considering states produced equally in all six dimensions.
1323: These states provide the needed pressure term
1324: at small $b$
1325: \bea \label{pressb}
1326: p^{(b)}_6&=& \frac{\tilde{n}_\chi(t)}{b} \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;
1327: \left( b<1 \right),
1328: \eea
1329: which will cause the motion of $b$ to return to the ESP.
1330:
1331: Given the additional sources, we now expect from the asymptotic behavior
1332: that the moduli should be trapped.
1333: However, the dynamics is actually quite involved due to the presence of non-linearities.
1334: Using the experimental result that $H_3>0$ (i.e. we live in three large dimensions),
1335: we examine the system numerically with the results appearing in Figures \ref{fig5} $-$ \ref{fig7}.
1336: In Fig. \ref{fig5}, we see the evolution of the radii $R_{11}$ and $b$ given initial conditions
1337: consistent with $H_3>0$.
1338: \begin{figure}
1339: \includegraphics[width=10.5cm]{fig5}
1340: \caption{\label{fig5} Evolution of $R_{11}$ (dark curve) and $b$ (light curve) for the exact density (\ref{rhogo}) and pressures, including membranes in all extra dimensions.}
1341: \end{figure}
1342: %
1343: \begin{figure}
1344: \includegraphics[width=10.5cm]{fig6}
1345: \caption{\label{fig6} $R_{11}$ for differing initial values of the expansion rate, with $\dot{R}_0=(0.001, 0.5)$ being barely distinguishable and with $\dot{R}_0= 0.6$ (light curve) we see the period begins to grow. We see that as the initial expansion rate is increased trapping takes longer to occur with $H_{11}>m_p$ leading to the case of no trapping.}
1346: \end{figure}
1347: %
1348: \begin{figure}
1349: \includegraphics[width=10.5cm]{fig7}
1350: \caption{\label{fig7} Evolution of $a(t)$ compared to $t^{1/2}$.}
1351: \end{figure}
1352: The jagged oscillations along the curve of $R_{11}$ are not due to numerical error,
1353: but rather
1354: result from the coupling to $b$ and the discontinuities associated with the pressure
1355: changing sign.
1356: We find that the radii will continue to oscillate with a decreasing amplitude, due to
1357: the expansion of $a(t)$.
1358: At late times we find that the radii $R_{11}$ and $b$ approach a constant value, which from (\ref{rhoapprox2}) implies
1359: $\rho \sim a^{-4}$ and $a(t) \sim t^{1/2}$.
1360: That is, our three dimensional universe evolves to that of a radiation dominated universe and
1361: the radii are trapped near the ESP $R_{11}=b=1$.
1362: We find that the trapping is robust, given that the initial expansion rates of the radii do
1363: not exceed the Planck scale (i.e., $H_6<m_p$ and $H_{11}<m_p$. This condition can be seen
1364: from Fig. \ref{fig6}, where we have plotted the evolution of $R_{11}$ for increasing values of
1365: the initial expansion rate $H_{11}$, and a similar result follows for $b(t)$.
1366: In Fig. \ref{fig7}
1367: we have presented a comparison of the late time behavior of $a(t)$ versus that of a radiation
1368: dominated universe $a(t)\sim t^{1/2}$. The wiggles in the evolution of $a(t)$ before the radii
1369: completely stabilize, naively may suggest the possibility of cosmological signatures coming from
1370: the trapping mechanism.
1371:
1372: \section{Conclusions and Future Prospects}
1373:
1374: We have considered both classical and quantum corrections to
1375: low-energy M-theory coming from dynamics of BPS membrane bound
1376: states whose effective mass depends on the radii of the extra
1377: dimensions. Including such states classically leads to an
1378: attractor mechanism that fixes moduli but is inconsistent with the
1379: use of the effective field theory approach.
1380:
1381: Insisting on an effective field theory description, we then
1382: consider quantum mechanical production of these membranes in a
1383: time-dependent background, and find the expected result that the
1384: production suffers Planckian suppression. However, we do find the
1385: possibility of significant and finite production for states that
1386: exhibit enhanced symmetry. We believe that these should correspond
1387: to non-threshold bound states of membranes and gravitational
1388: waves, which become tensionless at the eleven-dimensional
1389: self-dual point, $R_{11}=l_p$. An exact construction of such
1390: states is challenging, due to the problems of quantizing the
1391: supermembrane and the lack of an effective description of the
1392: theory in this regime.
1393:
1394: In \cite{Russo:1996if} it was shown that one can obtain the BPS
1395: spectrum of the supermembrane in special limits and by utilizing
1396: properties of BPS configurations. However, the enhanced states we
1397: are interested in require non-vanishing vacuum energy, which is
1398: not compatible with the $\mathcal{N}=2$ SUSY case considered in
1399: \cite{Russo:1996if}. Therefore, guided by the analysis of
1400: \cite{Russo:1996if}, we conjecture the existence of the enhanced
1401: states, awaiting a more concrete construction. One possibility for
1402: their existence is the case of $\mathcal{N}=1$ heterotic M-theory,
1403: in analogy with the enhanced gauge symmetry of heterotic strings.
1404: Another intriguing possibility is the recent conjecture that the
1405: spectrum of string / M-theory contains states that lie below the
1406: BPS bound \cite{Arkani-Hamed:2006dz} .
1407:
1408: Assuming that such enhanced states exist, we find that they can
1409: have a critical impact on the evolution of moduli, which would
1410: have been missed in the naive low-energy effective theory
1411: neglecting dynamics. We have found that, by including the
1412: backreaction of the membrane states produced, the radii are
1413: dynamically attracted to values near fixed points of S- and T-
1414: duality. This effect would be missed if one did not have a
1415: knowledge of (enhanced) M-theory states having masses which depend
1416: on evolving moduli. Thus, the lesson we learn is that the presence
1417: of time-dependence introduces a dynamical mass scale that must be
1418: taken into careful consideration in the effective field theory.
1419: Furthermore, it is paramount not to forget the string theory
1420: origin of the low-energy effective action.
1421:
1422: In addition to the corrections from on-shell particle production
1423: which we have explored here, there will be radiative corrections
1424: coming from the presence of the light membrane states. In fact, it
1425: was shown in \cite{Silverstein:2003hf} that, for the case of
1426: colliding D-branes, open strings becoming light lead to
1427: corrections to the gauge theory propagator resulting in a speed
1428: limit for the moduli. In the case we have considered here a
1429: similar story should hold, but this time it is the closed string
1430: moduli which should be slowing down (the radii of the extra
1431: dimensions). This offers a challenge, since the gauge theory
1432: interpretation of such processes is unclear. We find the
1433: possibility of speed limits for radii intriguing, and we hope to
1434: report on it shortly.
1435:
1436:
1437: \begin{acknowledgments}
1438: We would like to thank R. Brandenberger, B. Holdom, L. Kofman, A. Krause, D. Lowe, L. McAllister,
1439: S. Patil, D. Podolsky, S. Prokushkin, and H. Nastase for useful discussions. We would especially like to
1440: thank A. Jevicki, J. Russo, and A. Tseytlin for critical comments and suggestions and we are
1441: grateful to K. Hori for finding an error in an earlier draft. S.C. would like to thank the
1442: University of Toronto for hospitality. This work was financially supported in part by the
1443: National Science and Engineering Research Council of Canada, the Schnabel Woods institute,
1444: and the U.S. Department of Energy under contract DE-FG02-91ER40688, TASK A.
1445: \end{acknowledgments}
1446: \newpage
1447: \begin{thebibliography}{99}
1448:
1449: \bibitem{Vafa:2005ui}
1450: C.~Vafa,
1451: ``The string landscape and the swampland,''
1452: arXiv:hep-th/0509212.
1453: %%CITATION = HEP-TH 0509212;%%
1454:
1455: \bibitem{Arkani-Hamed:2006dz}
1456: N.~Arkani-Hamed, L.~Motl, A.~Nicolis and C.~Vafa,
1457: ``The string landscape, black holes and gravity as the weakest force,''
1458: arXiv:hep-th/0601001.
1459: %%CITATION = HEP-TH 0601001;%%
1460:
1461: \bibitem{Silverstein:2003hf}
1462: E.~Silverstein and D.~Tong,
1463: ``Scalar speed limits and cosmology: Acceleration from D-cceleration,''
1464: Phys.\ Rev.\ D {\bf 70}, 103505 (2004)
1465: [arXiv:hep-th/0310221].
1466: %%CITATION = HEP-TH 0310221;%%
1467:
1468: \bibitem{Kofman:2004yc}
1469: L.~Kofman, A.~Linde, X.~Liu, A.~Maloney, L.~McAllister and E.~Silverstein,
1470: ``Beauty is attractive: Moduli trapping at enhanced symmetry points,''
1471: JHEP {\bf 0405}, 030 (2004)
1472: [arXiv:hep-th/0403001].
1473: %%CITATION = HEP-TH 0403001;%%
1474:
1475: \bibitem{Watson:2004aq}
1476: S.~Watson,
1477: ``Moduli stabilization with the string Higgs effect,''
1478: Phys.\ Rev.\ D {\bf 70}, 066005 (2004)
1479: [arXiv:hep-th/0404177].
1480: %%CITATION = HEP-TH 0404177;%%
1481:
1482: \bibitem{Alishahiha:2004eh}
1483: M.~Alishahiha, E.~Silverstein and D.~Tong,
1484: ``DBI in the sky,''
1485: Phys.\ Rev.\ D {\bf 70}, 123505 (2004)
1486: [arXiv:hep-th/0404084].
1487: %%CITATION = HEP-TH 0404084;%%
1488:
1489: \bibitem{Brustein:2002xf}
1490: R.~Brustein, S.~P.~de Alwis and E.~G.~Novak,
1491: ``M-theory moduli space and cosmology,''
1492: Phys.\ Rev.\ D {\bf 68}, 043507 (2003)
1493: [arXiv:hep-th/0212344].
1494: %%CITATION = HEP-TH 0212344;%%
1495:
1496: \bibitem{Dine:1998qr}
1497: M.~Dine, Y.~Nir and Y.~Shadmi,
1498: ``Enhanced symmetries and the ground state of string theory,''
1499: Phys.\ Lett.\ B {\bf 438}, 61 (1998)
1500: [arXiv:hep-th/9806124].
1501: %%CITATION = HEP-TH 9806124;%%
1502:
1503: \bibitem{Russo:1996if}
1504: J.~G.~Russo and A.~A.~Tseytlin,
1505: ``Waves, boosted branes and BPS states in M-theory,''
1506: Nucl.\ Phys.\ B {\bf 490}, 121 (1997)
1507: [arXiv:hep-th/9611047].
1508: %%CITATION = HEP-TH 9611047;%%
1509:
1510: \bibitem{Schwarz:1995du}
1511: J.~H.~Schwarz,
1512: ``Superstring dualities,''
1513: Nucl.\ Phys.\ Proc.\ Suppl.\ {\bf 49}, 183 (1996)
1514: [arXiv:hep-th/9509148].
1515: %%CITATION = HEP-TH 9509148;%%
1516:
1517: \bibitem{Schwarz:1996bh}
1518: J.~H.~Schwarz,
1519: ``Lectures on superstring and M theory dualities,''
1520: Nucl.\ Phys.\ Proc.\ Suppl.\ {\bf 55B}, 1 (1997)
1521: [arXiv:hep-th/9607201].
1522: %%CITATION = HEP-TH 9607201;%%
1523:
1524: \bibitem{Bergshoeff:1987qx}
1525: E.~Bergshoeff, E.~Sezgin and P.~K.~Townsend,
1526: ``Properties Of The Eleven-Dimensional Super Membrane Theory,''
1527: Annals Phys.\ {\bf 185}, 330 (1988).
1528: %%CITATION = APNYA,185,330;%%
1529:
1530: \bibitem{Duff:1987cs}
1531: M.~J.~Duff, T.~Inami, C.~N.~Pope, E.~Sezgin and K.~S.~Stelle,
1532: ``Semiclassical Quantization Of The Supermembrane,''
1533: Nucl.\ Phys.\ B {\bf 297}, 515 (1988).
1534: %%CITATION = NUPHA,B297,515;%%
1535:
1536: \bibitem{Russo:1996ph}
1537: J.~G.~Russo,
1538: ``Supermembrane dynamics from multiple interacting strings,''
1539: Nucl.\ Phys.\ B {\bf 492}, 205 (1997)
1540: [arXiv:hep-th/9610018].
1541: %%CITATION = HEP-TH 9610018;%%
1542:
1543: \bibitem{Friess:2004zk}
1544: J.~J.~Friess, S.~S.~Gubser and I.~Mitra,
1545: ``String creation in cosmologies with a varying dilaton,''
1546: Nucl.\ Phys.\ B {\bf 689}, 243 (2004)
1547: [arXiv:hep-th/0402156].
1548: %%CITATION = HEP-TH 0402156;%%
1549:
1550: \bibitem{Battefeld:2005av}
1551: T.~Battefeld and S.~Watson,
1552: ``String gas cosmology,''
1553: arXiv:hep-th/0510022.
1554: %%CITATION = HEP-TH 0510022;%%
1555:
1556: \bibitem{Brandenberger:2005nz}
1557: R.~H.~Brandenberger,
1558: ``Challenges for string gas cosmology,''
1559: arXiv:hep-th/0509099.
1560: %%CITATION = HEP-TH 0509099;%%
1561:
1562: \bibitem{Brandenberger:2005fb}
1563: R.~H.~Brandenberger,
1564: ``Moduli stabilization in string gas cosmology,''
1565: arXiv:hep-th/0509159.
1566: %%CITATION = HEP-TH 0509159;%%
1567:
1568: \bibitem{Patil:2005nm}
1569: S.~P.~Patil,
1570: ``Moduli (dilaton, volume and shape) stabilization via massless F and D
1571: string modes,''
1572: arXiv:hep-th/0504145.
1573: %%CITATION = HEP-TH 0504145;%%
1574: %%Cited 14 times in SPIRES-HEP
1575:
1576: \bibitem{Kaloper:2004yj}
1577: N.~Kaloper, J.~Rahmfeld and L.~Sorbo,
1578: ``Moduli entrapment with primordial black holes,''
1579: Phys.\ Lett.\ B {\bf 606}, 234 (2005)
1580: [arXiv:hep-th/0409226].
1581: %%CITATION = HEP-TH 0409226;%%
1582:
1583: \bibitem{Abel:2005jx}
1584: S.~A.~Abel and J.~Gray,
1585: ``On the chaos of D-brane phase transitions,''
1586: JHEP {\bf 0511}, 018 (2005)
1587: [arXiv:hep-th/0504170].
1588: %%CITATION = HEP-TH 0504170;%%
1589:
1590: \bibitem{Mohaupt:2005pa}
1591: T.~Mohaupt and F.~Saueressig,
1592: ``Conifold cosmologies in IIA string theory,''
1593: Fortsch.\ Phys.\ {\bf 53}, 522 (2005)
1594: [arXiv:hep-th/0501164].
1595: %%CITATION = HEP-TH 0501164;%%
1596:
1597: \bibitem{Berndsen:2005qq}
1598: A.~Berndsen, T.~Biswas and J.~M.~Cline,
1599: ``Moduli stabilization in brane gas cosmology with superpotentials,''
1600: JCAP {\bf 0508}, 012 (2005)
1601: [arXiv:hep-th/0505151].
1602: %%CITATION = HEP-TH 0505151;%%
1603:
1604: \bibitem{Strominger:1995cz}
1605: A.~Strominger,
1606: ``Massless black holes and conifolds in string theory,''
1607: Nucl.\ Phys.\ B {\bf 451}, 96 (1995)
1608: [arXiv:hep-th/9504090].
1609: %%CITATION = HEP-TH 9504090;%%
1610:
1611: \bibitem{Lidsey:1999mc}
1612: J.~E.~Lidsey, D.~Wands and E.~J.~Copeland,
1613: ``Superstring cosmology,''
1614: Phys.\ Rept.\ {\bf 337}, 343 (2000)
1615: [arXiv:hep-th/9909061].
1616: %%CITATION = HEP-TH 9909061;%%
1617:
1618: \bibitem{Lawrence:1995ct}
1619: A.~E.~Lawrence and E.~J.~Martinec,
1620: ``String field theory in curved spacetime and the resolution of spacelike
1621: singularities,''
1622: Class.\ Quant.\ Grav.\ {\bf 13}, 63 (1996)
1623: [arXiv:hep-th/9509149].
1624: %%CITATION = HEP-TH 9509149;%%
1625:
1626: \bibitem{Gubser:2003vk}
1627: S.~S.~Gubser,
1628: ``String production at the level of effective field theory,''
1629: Phys.\ Rev.\ D {\bf 69}, 123507 (2004)
1630: [arXiv:hep-th/0305099].
1631: %%CITATION = HEP-TH 0305099;%%
1632:
1633: \bibitem{Birrell:1982ix}
1634: N.~D.~Birrell and P.~C.~W.~Davies,
1635: ``Quantum Fields In Curved Space,''
1636: %\href{http://www.slac.stanford.edu/spires/find/hep/www?irn=998621}{SPIRES entry}
1637:
1638: \bibitem{Chung:1998bt}
1639: D.~J.~H.~Chung,
1640: ``Classical inflation field induced creation of superheavy dark matter,''
1641: Phys.\ Rev.\ D {\bf 67}, 083514 (2003)
1642: [arXiv:hep-ph/9809489].
1643: %%CITATION = HEP-PH 9809489;%%
1644:
1645: \bibitem{Watson:2003gf}
1646: S.~Watson and R.~Brandenberger,
1647: ``Stabilization of extra dimensions at tree level,''
1648: JCAP {\bf 0311}, 008 (2003)
1649: [arXiv:hep-th/0307044].
1650: %%CITATION = HEP-TH 0307044;%%
1651:
1652: \end{thebibliography}
1653:
1654: \end{document}
1655:
1656: