1: \documentclass[aps,manuscript]{revtex4} \usepackage{graphics}
2: \input epsf
3: \def\ba{\begin{eqnarray}}
4: \def\ea{\end{eqnarray}}
5: \def\be{\begin{equation}}
6: \def\ee{\end{equation}}
7:
8: \def\VEV#1{\left\langle #1\right\rangle}
9: \def\mxth{\mathsurround=0pt }
10: \def\xversim#1#2{\lower2.pt\vbox{\baselineskip0pt \lineskip-.2pt
11: \ialign{$\mxth#1\hfil##\hfil$\crcr#2\crcr\sim\crcr}}}
12: \def\ltsim{\mathrel{\mathpalette\xversim <}}
13: \def\gtsim{\mathrel{\mathpalette\xversim >}}
14: \def\lsim{\mathrel{\mathpalette\xversim <}}
15: \def\gsim{\mathrel{\mathpalette\xversim >}}
16: \def\lta{\mathrel{\mathpalette\xversim <}}
17: \def\gta{\mathrel{\mathpalette\xversim >}}
18:
19: \newcommand{\labeq}[1] {\label{eq:#1}}
20: \newcommand{\eqn}[1] {(\ref{eq:#1})}
21: \newcommand{\labfig}[1] {\label{fig:#1}}
22: \newcommand{\fig}[1] {\ref{fig:#1}}
23: %\tighten
24: \begin{document}
25:
26: \title{M Theory Model of a Big Crunch/Big Bang Transition}
27:
28: \author{
29: Neil Turok$^1$, Malcolm Perry$^1$, and Paul J. Steinhardt$^2$}
30:
31: \affiliation{~}
32:
33: \affiliation{$^1$DAMTP, Centre for Mathematical Sciences,
34: Wilberforce Road, Cambridge CB3 0WA, UK }
35: \affiliation{$^2$Joseph Henry Laboratories,
36: Princeton University,
37: Princeton, NJ 08544, USA,\\
38: Institute for Advanced Studies, Olden Lane, Princeton, NJ 08540, USA.}
39:
40:
41: \begin{abstract}
42: We consider a picture in which the transition from a big crunch to a big bang
43: corresponds to the collision of two empty orbifold planes approaching each
44: other at a constant non-relativistic speed in a locally flat
45: background space-time, a situation relevant to
46: recently proposed cosmological
47: models. We show that $p$-brane states which wind around the extra dimension
48: propagate smoothly and unambiguously across the orbifold plane collision.
49: In particular we calculate the quantum mechanical production of winding
50: M2-branes extending from one orbifold to the other. We find
51: that the resulting density is finite and that the resulting
52: gravitational back-reaction is
53: small. These winding states, which include the string theory
54: graviton, can be propagated
55: smoothly across the transition using a perturbative expansion
56: in the membrane
57: tension, an expansion which from the point of view of string theory
58: is an expansion in
59: {\it inverse} powers of $\alpha'$. The conventional
60: description of a crunch based on Einstein general relativity,
61: involving Kasner or mixmaster behavior is misleading, we
62: argue, because general relativity is only the leading order
63: approximation to string theory in an expansion in positive
64: powers of $\alpha'$. In contrast, in the M theory setup we
65: argue that interactions should be well-behaved because of
66: the smooth evolution of the fields combined with the fact
67: that the string coupling tends to zero at the crunch.
68: The production of massive Kaluza-Klein states should also be
69: exponentially
70: suppressed for small collision speeds.
71: We contrast this good behavior with that found in
72: previous studies of strings in Lorentzian orbifolds.
73: \hfill\break
74: PACS number(s):
75: 11.25.-w,04.50.+h, 98.80.Cq,98.80.-k
76: \end{abstract}
77:
78: \maketitle
79:
80: %98.80.Cq inflation
81: %04.50.+h Gravity in more than four-dimensions
82: %98.80.-k cosmology in general
83: %11.25.-w Theory of fundamental strings
84: %
85: %
86:
87: \renewcommand\baselinestretch{1.0}
88: \tableofcontents
89:
90: \renewcommand\baselinestretch{1.0}
91:
92: \section{Introduction}
93:
94: One of the greatest challenges faced by string and M theory is
95: that of describing time-dependent singularities, such as
96: occur in cosmology and in black holes. These singularities
97: signal the catastrophic failure of general relativity at short
98: distances, precisely the pathology that string theory
99: is supposed to cure. Indeed string theory does succeed in
100: removing the divergences present in perturbative quantum gravity
101: about flat spacetime. String theory is also known to tolerate
102: singularities in certain static backgrounds
103: such as orbifolds
104: and conifolds.
105: However, studies within string theory thus far
106: have been unable to
107: shed much light on the far more interesting question of
108: the physical resolution of time-dependent singularities.
109:
110: In this paper we discuss M theory in one of the
111: simplest possible time-dependent backgrounds~\cite{kosst,steif},
112: a direct product of $d-1$-dimensional
113: flat Euclidean space $R^{d-1}$ with two dimensional
114: compactified Milne space-time, ${\cal M}_C$,
115: with
116: line element
117: \be
118: -dt^2+t^2 d\theta^2.
119: \labeq{line}
120: \ee
121: The compactified coordinate $\theta$
122: runs from $0$ to $\theta_0$.
123: As $t$ runs from $-\infty$ to $+\infty$, the
124: compact dimension shrinks away and reappears once
125: more, with
126: rapidity $\theta_0$.
127: Analyticity in $t$ suggests that this continuation
128: is unique~\cite{tolley1}.
129:
130: Away from $t=0$,
131: ${\cal M}_C$ is locally flat, as can be
132: seen by changing to coordinates $T=t\cosh \theta$, $Y=t\sinh \theta$
133: in which (\ref{eq:line}) is just $-dT^2+dY^2$. Hence
134: ${\cal M}_C\times R^{d-1}$ is naturally a
135: solution of any geometrical theory
136: whose field equations are
137: built from
138: the curvature tensor.
139: However, ${\cal M}_C\times R^{d-1}$ is nonetheless
140: mathematically singular at $t=0$
141: because the metric
142: degenerates when the compact dimension disappears.
143: General relativity cannot
144: make sense of this situation since
145: there ceases to be enough Cauchy data to
146: determine the future evolution
147: of fields. In fact, the situation is worse than this:
148: within general relativity,
149: generic perturbations diverge as log$|t|$ as
150: one approaches the singularity~\cite{ekperts},
151: signaling the breakdown
152: of perturbation theory and the approach to
153: Kasner or mixmaster behavior, according to which the
154: space-time
155: curvature diverges as $t^{-2}$.
156: Of course, this breakdown of general
157: relativity presents a challenge:
158: can M theory make sense of the singularity at $t=0$?
159:
160: We are interested in
161: what happens in the immediate
162: vicinity of $t=0$, when the compact dimension
163: approaches, and becomes smaller than,
164: the fundamental membrane tension scale.
165: The key
166: difference between M theory (or string theory) and local field theories
167: such as general relativity is the existence of extended
168: objects including those stretching across
169: compactified dimensions.
170: Such states become very light as the compact
171: dimensions shrink below the fundamental scale.
172: They are known to play a central
173: role in resolving
174: singularities for example in orbifolds and in
175: topology-changing transitions~\cite{greene}. Therefore, it
176: is very natural to ask what role such states
177: play in big crunch/big bang space-times.
178:
179: In this paper we shall show
180: that
181: $p$-branes winding uniformly around the compact
182: dimension
183: obey equations, obtained by canonical methods,
184: which are completely
185: regular at $t=0$. These methods are naturally
186: invariant under choices of worldvolume coordinates.
187: Therefore we claim that it is possible to
188: unambiguously describe evolution of such states
189: from $t<0$ to $t>0$, through a cosmological
190: singularity from the point of view of the low energy
191: effective theory.
192: Indeed the space-time
193: we consider corresponds locally to one where
194: two empty, flat, parallel orbifold planes collide, precisely
195: the situation envisaged in recently proposed cosmological
196: models.
197:
198: Hence, the calculations we report are directly relevant to the
199: ekpyrotic~\cite{kost} and
200: cyclic universe~\cite{STu} scenarios, in which passage
201: through a singularity of this type is taken
202: to represent the standard hot big bang.
203: In particular, the equation of state during the dark energy and
204: contracting phases causes the orbifold planes to be empty,
205: flat, and parallel as they approach to within a string
206: length~\cite{chaos,review}.
207: This setup makes it natural to split the study
208: of the collision of orbifold planes into a separate analysis of the
209: winding modes, which become light near $t=0$, and
210: other modes that become heavy there. This strategy feeds directly into
211: the considerations in this paper.
212:
213: The physical reason why winding states are well-behaved is easy to
214: understand. The obvious problem with a space-time such as
215: compactified Milne is the blue shifting effect felt by
216: particles which can
217: run around the compact dimension as it shrinks away.
218: As we shall discuss in detail, winding
219: states wrapping around the compact dimension
220: do not feel any blue shifting effect because there
221: is no physical motion
222: along their length. Instead, as their length disappears,
223: from the point of view of the noncompact dimensions,
224: their effective mass or tension
225: tends to zero but their energy and momentum remain finite. When
226: such states are quantized the corresponding fields are well-behaved
227: and the field equations are analytic at $t=0$.
228: In contrast, for bulk, non-winding states, the
229: motion in the $\theta$ direction is physical and it becomes
230: singular as $t$ tends
231: to zero. In the quantum field theory of such states,
232: this behavior results in
233: logarithmic divergences of the
234: fields near $t=0$, even for the lowest
235: modes of the field which
236: are uniform in $\theta$ {\it i.e.,} the lowest Kaluza-Klein
237: modes (see Section X and Appendix 4).
238:
239: We are specially interested in the case of M theory,
240: considered as the theory of branes. As
241: the compact dimension becomes small, the
242: winding M2-branes we focus on
243: are the lowest energy states of the theory, and describe
244: a string theory in a certain time-dependent background.
245: The most remarkable feature of this setup is that the
246: string theory includes a graviton and, hence,
247: describes perturbative gravity near $t=0$.
248: In this paper, we show these strings,
249: when considered as winding M2-branes,
250: follow smooth evolution (see Section VII) across the singularity,
251: even though the string frame metric degenerates there.
252: Furthermore we show that this good behavior is only
253: seen in a perturbation expansion in the membrane tension,
254: corresponding from the string theory point of view to an expansion
255: in {\it inverse powers of $\alpha'$}. We argue that
256: the two-dimensional
257: nonlinear sigma model describing this situation
258: is renormalizable in such an expansion.
259: The good behavior of the relevant string theory
260: contrasts sharply with the bad behavior of general
261: relativity. There is no contradiction, however,
262: because general relativity is only the first approximation
263: to string theory in an expansion in positive
264: powers of $\alpha'$. Such an expansion is valid
265: when $t$ is much larger than the fundamental membrane scale,
266: but it fails near the singularity where, as mentioned,
267: the theory is regular in the opposite $(\alpha')^{-1}$ expansion.
268: The logarithmic divergences of perturbations found
269: using the Einstein equations
270: are, thereby, seen to be due to the failure
271: of the $\alpha'$ expansion, and not of M or string theory {\it per se}.
272:
273: When the M theory dimension is small,
274: the modes of the theory are neatly partitioned
275: into light $\theta$-independent modes and heavy
276: $\theta$-dependent modes.
277: The former set consists of winding membranes, which
278: describe a string theory including perturbative gravity.
279: This is the sector
280: within which cosmological perturbations
281: lie, and which will be our prime focus in this paper.
282:
283: The $\theta$-dependent modes are likely to be harder to
284: describe.
285: The naive argument that these modes are problematic
286: because they are blue shifted and, hence, infinitely
287: amplified as $t\rightarrow 0$
288: is suspect because it
289: relies on conventional
290: Einstein gravity,
291: Here we argue that,
292: close to the brane collision, Einstein gravity
293: is a poor approximation and, instead, perturbative
294: gravity is described by
295: the non-singular winding sector. The latter
296: does not
297: exhibit blue shifting behavior near $t=0$,
298: so the naive argument does not apply.
299:
300: Witten
301: has argued \cite{witten} that
302: the massive Kaluza-Klein modes of the
303: eleven-dimensional
304: theory
305: map onto non-perturbative black hole states in the effective
306: string theory.
307: Even though these black hole states are likely to be
308: hard to describe in detail, we will explain in Section II
309: why their overall effect is likely to be small.
310: First, in the cosmological scenarios of interest,
311: the universe enters the regime where perturbative gravity is
312: described by the winding modes ({\it i.e.,} the branes are close)
313: with a negligible density
314: of Kaluza-Klein massive modes. This suppression is a result of
315: the special
316: equation of state in the contracting phase that precedes this
317: regime~\cite{review}. Second, the density of black holes quantum produced
318: due to the time-dependent background in the
319: vicinity of the collision is likely to be
320: negligible because they are so massive
321: and so large.
322:
323:
324: For these reasons, we focus at present on the propagation
325: of the perturbative gravity sector near $t=0$
326: corresponding to the winding M2-brane states.
327: In Section II, we introduce the compactified Milne background
328: metric that describes the collision between orbifold planes in the
329: big crunch/big bang transition. We also discuss the motivation for
330: the initial conditions that will be assumed in this paper.
331: The canonical Hamiltonian description of $p$-branes in
332: curved space is given in Section III and applied to winding modes
333: in the compactified Milne background in Section IV.
334: Section V discusses the key difference in the Hamiltonian
335: description between winding and bulk states that accounts for
336: their different behavior near the big crunch/big bang
337: transition.
338:
339: Section VI is the consideration of a toy model in which
340: winding strings are produced as the
341: branes collide. The winding modes
342: are described semi-classically, and their quantum
343: production at the bounce is computed. Section VII presents the
344: analogous semi-classical description of winding M2-branes.
345: Although we cannot solve the theory
346: exactly, we show the eleven-dimensional theory
347: is well behaved near $t=0$
348: and explain how the apparent singularity in the dimensionally-reduced
349: string theory is resolved in the membrane picture. Then, Section VIII
350: makes clear the difference between our calculation, an expansion in inverse
351: powers of $\alpha'$, versus Einstein gravity, the leading term in
352: an expansion
353: in positive powers of
354: $\alpha'$. This argument is key to explaining why we think the
355: transition is calculable even though it appears to be poorly
356: behaved when described by Einstein gravity.
357: Section IX, then,
358: uses
359: Euclidean instanton methods to study
360: the quantum production of
361: winding M2-branes (in analogy to the case of winding strings in Section VI)
362: induced by
363: passage through the singularity, obtaining
364: finite and physically sensible results. In particular,
365: the resultant density
366: tends to zero as the speed
367: of contraction of the compact dimension is reduced.
368: We estimate the gravitational back-reaction
369: and show it is small provided $\theta_0$,
370: the rapidity of contraction of the compact dimension,
371: is small.
372:
373: In Section X, we comment on why our M theory setup
374: is better behaved than the Lorentzian orbifold case~\cite{seibergetal}
375: considered in some previous investigations of the big crunch/big
376: bang transition. The fundamental problem with the latter case,
377: we argue, is that perturbative gravity lies within the
378: bulk sector and not the winding sector as far as
379: the compactified Milne singularity is concerned.
380: Therefore, it is susceptible to the blueshifting
381: problem mentioned above, rendering the string equations singular.
382:
383:
384: \section{The Background Big Crunch/Big Bang Space-time}
385:
386: The $d+1$-dimensional
387: space-time we consider is a direct product of
388: $d-1$-dimensional Euclidean space, $R^{d-1}$, and a
389: two-dimensional time-dependent space-time
390: known as compactified
391: Milne space-time, or ${\cal M}_C$.
392: The line element for ${\cal M}_C \times R^{d-1}$ is thus
393: \be
394: ds^2= -dt^2 +t^2 d\theta^2 + d\vec{x}\,^2, \qquad 0\leq \theta \leq \theta_0,
395: \qquad -\infty <t <\infty,
396: \labeq{backg}
397: \ee
398: where $\vec{x}$ are Euclidean
399: coordinates on $R^{d-1}$, $\theta$
400: parameterizes the compact dimension and $t$ is the time.
401: The compact dimension may
402: either be a circle, in which case we identify $\theta$
403: with $\theta+\theta_0$, or a $Z_2$ orbifold in which case we
404: identify $\theta$ with $\theta+2 \theta_0$ and
405: further identify $\theta$ with $2\theta_0-\theta$.
406: The fixed points $\theta=0$ and $\theta=\theta_0$ are
407: then interpreted as tensionless $Z_2$-branes approaching
408: at rapidity $\theta_0$, colliding at $t=0$ to re-emerge
409: with the same relative rapidity.
410:
411: The orbifold reduction is the case of prime interest in the
412: ekpyrotic/cyclic models, originally motivated by the
413: construction of heterotic M theory
414: from eleven dimensional supergravity~\cite{hw,ovrut}.
415: In these models, the boundary branes possess nonzero tension.
416: However, the tension is a subdominant effect near $t=0$ and
417: the brane collision is locally well-modeled by ${\cal M}_C \times R^{d-1}$
418: (See Ref.~\cite{tolleyperts}).
419:
420: The line element (\ref{eq:backg}) is of particular interest
421: because it is locally flat and, hence, an exact solution not only of
422: $d+1$-dimensional Einstein gravity but of
423: any higher dimensional gravity theory whose field equations
424: are constructed from curvature invariants with no cosmological
425: constant.
426: And even if a small cosmological constant
427: were present, it would not have a large effect locally
428: so that
429: solutions with a similar local structure
430: in the vicinity of the singularity
431: would be expected to exist.
432:
433: Consider the description of (\ref{eq:backg}) within
434: $d+1$-dimensional general relativity.
435: When the compact dimension is small, $\theta$-dependent
436: states become massive and it makes sense to describe the
437: system using a low energy effective field theory. This may be
438: obtained by the well known procedure of
439: dimensional reduction.
440: The $d+1$-dimensional line element (\ref{eq:backg}) may be rewritten
441: in terms of a $d$-dimensional Einstein frame metric, $g^{(d)}_{\mu \nu}$,
442: and a scalar field $\phi$:
443: \be
444: ds^2 = e^{2\phi \sqrt{(d-2)/(d-1)}} d\theta^2 +
445: e^{-2\phi /\sqrt{(d-2)(d-1)}}g^{(d)}_{\mu \nu} d x^{\mu} d x^{\nu}.
446: \labeq{metr}
447: \end{equation}
448: The numerical coefficients are chosen so that
449: if one substitutes
450: this metric into the $d+1$-dimensional Einstein
451: action and assumes that $\phi$ and $g^{(d)}_{\mu \nu}$ are both
452: $\theta$-independent, one obtains
453: $d$-dimensional Einstein gravity with
454: a canonically normalized
455: massless, minimally coupled scalar field $\phi$.
456: (Here we choose units in which the
457: coefficient of the Ricci scalar in the $d$-dimensional
458: Einstein action is ${1\over 2}$. We have also ignored Kaluza-Klein
459: vectors, which play no role in this argument and
460: are in any case projected out in the orbifold reduction.)
461:
462: From the viewpoint of the low energy effective theory, the
463: $d+1$-dimensional space-time ${\cal M}_C \times R^{d-1}$
464: is reinterpreted as a
465: $d$-dimensional cosmological solution where $t$ plays the
466: role of the conformal time.
467: Comparing (\ref{eq:backg}) and (\ref{eq:metr}),
468: the $d$-dimensional Einstein-frame metric
469: $g^{(d)}_{\mu \nu}=a^2\, \eta_{\mu \nu}$ with $a
470: \propto |t|^{1/(d-2)}$ and the scalar field $\phi =
471: \sqrt{ (d-1)/(d-2)} {\rm ln}|t|$. From this point of view
472: $t=0$ is a space-like curvature
473: singularity of the standard big bang type where
474: the scalar
475: field diverges, and
476: passing through
477: $t=0$ would seem to be impossible.
478: However, by lifting to the higher dimensional viewpoint
479: one sees that the
480: situation is not really so bad. The line element (\ref{eq:backg})
481: is in fact static at all times in the noncompact
482: directions $\vec{x}$. So for example, matter
483: localized on the branes would see no blue shifting effect
484: as the singularity approaches\cite{STu}. As we discuss
485: in detail in Section IV,
486: winding states do not see a blue shifting effect either.
487:
488: In this paper, we consider an M theory picture
489: with two empty, flat, parallel colliding orbifold planes
490: and we are interested in the dynamics of the collision region
491: from the point where the planes are roughly a string length apart.
492: The assumed initial conditions are important
493: for two reasons. First, they correspond to the simple compactified
494: Milne background discussed above.
495: Second, as mentioned in the introduction,
496: this initial condition means that the
497: excitations neatly divide into
498: light winding modes that are becoming massless
499: and heavy Kaluza-Klein modes that are becoming massive and decoupling
500: from the low-energy effective theory.
501:
502: What we want to show now is that initial conditions with negligible
503: heavy Kaluza-Klein modes present are naturally produced in cosmological
504: scenarios such a the cyclic model~\cite{STu,review}.
505: This justifies our focus
506: on the winding modes throughout the remainder of the paper.
507: However, the argument is inessential to the rest of the paper
508: and readers willing to
509: accept the initial conditions without justification may wish
510: to
511: proceed straight away to the next Section.
512:
513: The cyclic model assumes a non-perturbative
514: potential hat produces an attractive force between the orbifold
515: planes. When the branes are far apart, perhaps $10^4$ Planck lengths,
516: the potential energy is positive and small, acting as the dark energy
517: that causes the currently observed accelerated expansion.
518: In the dark energy dominated
519: phase, the branes stretch by a factor of two in linear dimensions
520: every 14 billion years or so, causing the branes to become flat, parallel
521: and empty. In the low-energy effective theory,
522: the total energy
523: is dominated by the scalar field $\phi$ whose value determines the
524: distance between branes.
525: As the planes draw together, the
526: potential energy $V(\phi)$ of this field
527: decreases and becomes increasingly negative until
528: the expansion stops and a contracting phase begins.
529:
530: A key point is that this contracting phase is described by
531: an attractor solution, which has an equation of state parameter
532: $w\equiv P/\rho \gg 1$. The energy density of the scalar field
533: $\phi$ scales as
534: \be
535: \rho_\phi \propto a^{-(d-1)(1+w)}
536: \labeq{rhophi}
537: \ee
538: in this phase. This is a very rapid increase, causing the density
539: in $\phi$
540: to come to dominate over curvature,
541: anisotropy, matter, or radiation\cite{chaos}.
542: We now show that $\phi$ comes to dominate over
543: the massive Kaluza-Klein modes. The latter
544: scale as
545: \be
546: \rho_{KK} \propto a^{-(d-1)} L^{-(d-1)/(d-2)}
547: \labeq{rhoKK}
548: \ee
549: where $L$ is the size of the extra dimension. The first factor
550: is the familiar inverse volume scaling which all particles suffer.
551: The second factor indicates the effective mass of the
552: Kaluza-Klein modes. The $d+1$-dimensional mass is
553: $L^{-1}$, but this must be converted to a $d-$dimensional
554: mass using the ratios of square roots of
555: the $00$ components of the $d+1$-dimensional
556: metric and the $d$-dimensional metric. This correction
557: produces the second factor in (\ref{eq:rhoKK}).
558:
559: From (\ref{eq:metr}),
560: we have $L\propto e^{\phi\sqrt{(d-2)/(d-1)}}$. Now the key
561: point is that this scales much more slowly with $a$ than
562: the potential $V(\phi)$ which scales as $\rho_\phi$ in
563: (\ref{eq:rhophi}). Neglecting the scaling with $L$,
564: the density of massive Kaluza-Klein modes scales as
565: as $\rho_\phi^{1/(1+w)}$. The final suppression of the
566: density of massive modes relative to the density in $\phi$ is
567: therefore $\sim (V_{i}/V_{f})^{w/(1+w)}$ where $V_{i}$ and $V_{f}$ are the
568: magnitudes of the scalar potential when the $w\gg 1$ phase
569: begins and ends. For large $w$, which we need in order to
570: obtain scale-invariant perturbations, this is
571: an exponentially
572: large factor~\cite{STu,review}.
573:
574: The massive Kaluza-Klein modes are, hence, exponentially
575: diluted when the $w\gg 1$ phase ends and the Milne
576: phase begins. During the Milne phase,
577: the scalar field is massless and has an equation of state
578: $w=1$, so $\rho_\phi$ scales as $a^{-2(d-1)}$ as
579: the distance between the orbifold planes
580: shrinks to zero.
581: In this regime, the Kaluza-Klein massive mode density
582: scales in precisely the same way. Therefore, their
583: density remains an exponentially small fraction of
584: the total density right up to collision: meaning from the
585: string theory point of view that the black
586: hole states remain exponentially rare.
587:
588: So we need only worry about
589: black holes produced in the vicinity of the
590: brane collision itself. From the point of view
591: of the higher dimensional theory, the oscillation
592: frequency of the masssive Kaluza-Klein modes
593: $\omega \sim |\theta_0t|^{-1}$
594: changes adiabatically,
595: $\dot{\omega}/\omega^2 \sim \theta_0 \ll 1$ for
596: small $\theta_0$, all the way to $t=0$. Therefore,
597: one expects little particle production before or
598: after $t=0$.
599: From the dimensionally reduced point of view, the
600: mass of the string theory black holes
601: is larger than the Hubble constant $\sim t^{-1}$,
602: by the same factor $\theta_0^{-1}$.
603: From either analysis, production
604: of such states should
605: be suppressed by
606: a factor $e^{-1/\theta_0}$, making it negligible
607: for small $\theta_0$.
608:
609: %PAUL
610: In sum, for the cosmological models of interest,
611: the Kaluza-Klein modes are
612: exponentially rare
613: when the Milne phase begins, and, since their mass increases
614: as the collision approaches, they should not be generated by the
615: orbifold plane motion.
616: (They effectively decouple from the low energy effective
617: theory.)
618: Hence, all the properties we want at the outset of our calculation here
619: are naturally achieved by the contracting phase with $w\gg1$, as occurs
620: in some current cosmological models\cite{review}.
621: %PAULend
622:
623:
624:
625:
626: \section{General Hamiltonian for $p$-Branes in Curved Space}
627:
628: The classical and quantum dynamics of $p$-branes may be
629: treated using canonical methods, indeed $p$-branes
630: provide an application {\it par
631: excellence} of Dirac's general method. As
632: Dirac himself
633: emphasized~\cite{dirac}, one of the advantages
634: of the canonical approach is that
635: it
636: allows a completely general choice of gauge.
637: In contrast, gauge
638: fixed methods tie one to a choice of gauge before it is
639: apparent whether that gauge is or isn't a good choice. In the
640: situation of interest here, where the background space-time
641: is singular,
642: the question of
643: gauge
644: choice is especially delicate. Hence, the canonical
645: approach is preferable.
646:
647: In this Section we provide an overview of the main results.
648: The technical details are relegated to Appendix 1. Our starting point is
649: the Polyakov action for a $p$-brane described by embedding
650: coordinates $x^\mu$ in a
651: a background space-time with metric
652: $g_{\mu \nu}$:
653: \be
654: {\cal S}_p = -{1\over 2} \mu_p \int d^{p+1} \sigma
655: \sqrt{-\gamma}\left(\gamma^{\alpha \beta}
656: \partial_\alpha x^\mu \partial_\beta x^\nu g_{\mu \nu} -(p-1) \right),
657: \labeq{pact}
658: \ee
659: where $\mu_p$ is a mass per unit $p$-volume. The $p$-brane worldvolume
660: has coordinates $\sigma^\alpha$, where $\sigma^0 =\tau$ is the
661: time and $\sigma^i$, $i=1..p$ are the spatial
662: coordinates.
663:
664: Variation of the action with respect to $\gamma_{\alpha \beta}$
665: yields the constraint that for $p\neq 1$,
666: $\gamma_{\alpha \beta}$ equals
667: the induced metric $\partial_\alpha x^\mu \partial_\beta x^\nu g_{\mu \nu}$
668: whereas for $p= 1$ $\gamma_{\alpha \beta}$ is conformal
669: to the induced metric. Substituting these results
670: back into the action one obtains the Nambu action for the
671: embedding coordinates $x^\mu(\sigma^\alpha)$ {\it i.e.,} $-\mu_p$ times the
672: induced $p$-brane world volume. We shall go back and forth between
673: the Polyakov and Nambu forms in this paper.
674: The former is preferable for quantization but the latter is still
675: useful for
676: discussing
677: classical solutions.
678:
679: The simplest case of (\ref{eq:pact}) is $p=0$,
680: a $0$-brane or massive particle. Writing
681: $\gamma_{00}=-e^2$ with $e$ the `einbein', one obtains
682: \be
683: {\cal S}_0 = {1\over 2} m\int d \tau
684: \left(e^{-1}
685: \dot{x}^\mu \dot{x}^\nu g_{\mu \nu} -e\right),
686: \labeq{parta}
687: \ee
688: where we have set $\mu_0=m$
689: and the dot above a variable indicates a derivative with respect to $\tau$.
690: Variation with respect to $e$ yields the constraint $e^2= -
691: \dot{x}^\mu \dot{x}^\nu g_{\mu \nu}$.
692: The canonical momentum is $p_\mu = m g_{\mu \nu}\dot{x}^\nu e^{-1}$ and
693: the constraint implies the familiar mass shell
694: condition $g^{\mu \nu} p_\mu p_\nu = -m^2$.
695:
696: %The next case, the $1$-brane or string, has the special
697: %feature that there is
698: %no cosmological constant term, and that
699: %the classical action is invariant
700: %under conformal transformations
701: %$\gamma_{\alpha \beta} \rightarrow \Omega^2 \gamma_{\alpha \beta}$.
702: %The next case, $2$-branes, are specially interesting because they
703: %and their duals, $5$-branes, are
704: %the fundamental entities in 11-dimensional M theory.
705: %Winding $2$-branes are of special interest
706: %because, for small compactification radius,
707: %they describe a string theory that includes
708: %perturbative gravity.
709: %Nevertheless
710: %much of our analysis will hold for general $p$
711: %so we shall
712: %frame it in those terms.
713:
714:
715: The canonical treatment for general $p$
716: is explained in
717: Appendix 1.
718: The main result is that a $p$-brane obeys $p+1$
719: constraints, reading
720: \be
721: C\equiv \pi_\mu \pi_\nu g^{\mu \nu}
722: +\mu_p^2 {\rm Det}( {x}^\mu_{,i} x^\nu_{,j} g_{\mu \nu})\approx 0, \qquad
723: C_i\equiv x^\mu_{,i} \pi_\mu\approx 0,
724: \labeq{consta}
725: \ee
726: where `$\approx 0$' means `weakly zero' in sense of the Dirac canonical
727: procedure (see Appendix I). Here the brane embedding coordinates
728: are $x^\mu$ and their conjugate momentum densities are
729: $\pi_\mu$.
730: The spatial worldvolume coordinates are
731: $\sigma^i$, $i=1,\dots,p$, and the
732: corresponding partial derivatives are
733: denoted $x^\mu_{,i}$.
734: The quantity
735: ${x}^\mu_{,i} x^\nu_{,j} g_{\mu \nu}$ is
736: the induced spatial metric on the $p$-brane. In Appendix 2
737: we calculate the Poisson bracket algebra of
738: the constraints (\ref{eq:consta}), showing that
739: the algebra closes and hence the constraints are all first class.
740: The constraints (\ref{eq:consta}) are invariant under worldvolume
741: coordinate transformations.
742:
743: The Hamiltonian giving the most general
744: evolution in worldvolume time $\tau$ is then given by
745: \be
746: H=\int d^p \sigma \left({1\over 2} A C + A^i C_i\right),
747: \labeq{totalh1}
748: \ee
749: with $C$ and $C_i$ given in (\ref{eq:consta}).
750: The functions $A$ and $A^{i}$ are
751: completely arbitrary, reflecting the
752: arbitrariness in the choice of worldvolume time and space
753: coordinates. All coordinate choices related by nonsingular
754: coordinate transformations give equivalent physical
755: results.
756:
757: For $p=0$, anything
758: with a spatial index $i$ can be ignored, except
759: the determinant in (\ref{eq:consta}) which is replaced by
760: unity. The first constraint is then the usual mass shell
761: condition, and the Hamiltonian is an
762: arbitrary function
763: of $\tau$ times the constraint.
764: The case of $p=1$, {\it i.e.,} a string, in Minkowski space-time,
765: $g_{\mu \nu} = \eta_{\mu \nu}$ is also simple and familiar.
766: In this case, the
767: constraints and the Hamiltonian (\ref{eq:totalh1}) are quadratic.
768: The resulting equations of motion are linear and hence
769: exactly
770: solvable. The
771: constraints (\ref{eq:consta}) amount
772: to the usual Virasoro conditions.
773: In general, the $p+1$ constraints (\ref{eq:consta}) together
774: with the $p+1$ free choices
775: of gauge functions $A$ and $A_i$ reduce
776: the number of physical coordinates and momenta
777: to $2(d+1) - 2(p+1)= 2(d-p)$, the correct number of transverse
778: degrees of freedom for a $p$-brane in $d$ spatial dimensions.
779:
780:
781:
782: \section{Winding $p$-Branes in ${\cal M}_C \times R^{d-1}$}
783:
784: In this paper, we shall study the dynamics of branes which wind
785: around the compact dimension in ${\cal M}_C \times R^{d-1}$,
786: the line element for which is given in (\ref{eq:backg}).
787: This space-time possesses an isometry
788: $\theta \rightarrow \theta +$constant, so one
789: can consistently truncate the theory to consider
790: $p$-branes which
791: wind uniformly around the $\theta$ direction.
792: Such configurations may be described by
793: identifying one of the $p$-brane spatial
794: coordinates (the $p$'th spatial coordinate,
795: $\sigma^p$ say) with $\theta$ and to simultaneously
796: insist that
797: that $\partial_p x^\mu =\partial_p \pi_\mu=0$.
798:
799: Through Hamilton's equations, the constraint
800: $\theta=\sigma^p$ implies
801: that $\pi_\theta=0$. This suggests that we can set
802: $\theta=\sigma^p$
803: and $\pi_\theta=0$ and, hence, dimensionally reduce the $p$-brane
804: to a $(p-1)$-brane. Detailed confirmation that this
805: is indeed consistent proceeds as follows.
806: We compute the Poisson brackets between all the
807: constraints $C$, $C_i$, $\theta-\sigma^p$ and $\pi_\theta$.
808: Following the Dirac procedure, we then
809: attempt
810: to build a maximal set of first class constraints.
811: The constraints $C$ and $C_i$ commute with each other,
812: for all $\sigma^i$, but not with
813: $\theta-\sigma^p$ and $\pi_\theta$.
814: The solution is to remove all
815: the $\pi_\theta$ and $\theta_{,p}$ terms from $C$ and $C_\theta$
816: by adding terms involving
817: $\pi_\theta$ and $\theta_{,p} -1 = (\theta-\sigma^p)_{,p}$.
818: The new $C$ and $C_i$ are now
819: first class since they have weakly vanishing Poisson brackets with all
820: the constraints, and the remaining second class constraints are
821: $\theta-\sigma^p$ and $\pi_\theta$. For these,
822: construction of the Dirac bracket is trivial
823: and it amounts simply to canceling the $\theta$ and
824: $\pi_\theta$ derivatives from the Poisson bracket.
825: The conclusion is that we can indeed consistently
826: set $\theta=\sigma^p$
827: and $\pi_\theta=0$.
828: We shall see in the following section that eliminating
829: $\theta$ and $\pi_{\theta}$ in this way results directly
830: in the
831: good behavior of the winding modes as $t \rightarrow 0$,
832: in contrast with the bad behavior of bulk modes.
833:
834: The surviving first class constraints for winding $p$-branes are
835: those obtained by substituting $\theta=\sigma^p$ and
836: $\pi_\theta=0$ into the $p$-brane constraints (\ref{eq:consta}),
837: namely
838: \be
839: C\equiv \pi_\mu \pi_\nu \eta^{\mu \nu}
840: +\mu_{p}^2 \theta_0^2 t^2 {\rm Det}( {x}^\mu_{,i} x^\nu_{,j}
841: \eta_{\mu \nu})\approx 0;\qquad
842: C_i\equiv x^\mu_{,i} \pi_\mu\approx 0,
843: \labeq{redcon}
844: \ee
845: where $i$ and $j$ now run from $1$ to $p-1$ and $\mu$ and
846: $\nu$ from $0$ to $d$. The $t^2$ term comes from the
847: $\theta \theta$ component of the ${\cal M}_C \times R^{d-1}$
848: background metric (\ref{eq:backg}).
849: We have also re-defined the
850: momentum density $\pi_\mu$ for the $p-1$-brane to be $\theta_0$
851: times the momentum density for the $p$-brane
852: so that the new Poisson brackets are correctly
853: normalized to give a $p-1$-dimensional delta function.
854: The Hamiltonian is again given by the form (\ref{eq:totalh1})
855: with the integral taken over the remaining $p-1$ spatial
856: coordinates.
857:
858: For $p=1$, the
859: reduced string is a $d$-dimensional particle.
860: $\pi_\mu$ is now the momentum $p_\mu$
861: and the determinant appearing
862: in (\ref{eq:redcon}) should be interpreted as
863: unity. The second constraint is
864: trivial since there are no remaining spatial
865: directions. The general Hamiltonian reads:
866: \be
867: H_0= A(\tau) \left( p_\mu p_\nu \eta^{\mu \nu}
868: +\mu_{1}^2 \theta_0^2 t^2 \right),
869: \labeq{hemp}
870: \ee
871: where $\mu, \nu$ run from $0$ to $d-1$ and
872: $A(\tau)$ is an arbitrary function of $\tau$. We shall
873: study the quantum field theory for this Hamiltonian in
874: Section VI.
875:
876: %For $p=2$ the determinant reduces to
877: %${x^\mu}' {x^\nu}' g_{\mu \nu}$, where prime is the derivative
878: %with respect to the worldsheet coordinate $\sigma$,
879: %and the
880: %second constraint implies the momentum density must
881: %be normal to the length of the 1-brane {\it i.e.,} string.
882: %In general, the reduced $p-1$-brane possesses
883: % $p$-dimensional coordinate invariance
884: %and the corresponding general Hamiltonian
885: %is the $p-1$ dimensional
886: %integral of an arbitrary linear combination
887: %of the two constraints in (\ref{eq:redcon}).
888:
889: Comparing
890: (\ref{eq:redcon}) with (\ref{eq:consta}), we see that
891: a $p$-brane which winds around the compact
892: dimension in ${\cal M}_C \times R^{d-1}$ behaves like
893: a $p-1$-brane in Minkowski spacetime with a time-dependent
894: effective
895: tension $\mu_p \theta_0 |t|$, {\it i.e.,} the $p$-brane
896: tension times the size of the compact dimension, $\theta_0 |t|$.
897:
898:
899: \section{Winding states versus bulk states}
900: \label{bulk}
901:
902: We have discussed in detail how in the canonical treatment
903: the coordinate $\theta$ and conjugate momentum density
904: $\pi_\theta$
905: may be eliminated
906: for $p$-branes winding
907: uniformly around the compact dimension.
908: This is physically reasonable, since
909: motion of
910: a winding $p$-brane along its own length (i.e. along $\theta$)
911: is meaningless. This is a crucial
912: difference from bulk states. Whereas the metric
913: on the space of coordinates for bulk states includes
914: the $t^2 d\theta^2$ term, the metric on the
915: space of coordinates for winding states does not.
916: As we discuss in detail in Appendix 4, when we quantize
917: the system the square root of the
918: determinant of the metric on the space
919: of coordinates appears in the quantum field
920: Hamiltonian. For bulk modes the
921: metric on the space of coordinates
922: inherits the singular behavior of the background
923: metric
924: (\ref{eq:backg}), degenerating at $t=0$ so that
925: causing the field equations to
926: become singular at $t=0$ even for $\theta$-independent
927: field modes (see Section X). Conversely, for winding modes the Hamiltonian
928: operator is regular at $t=0$.
929:
930: The metric on the space of coordinates is defined
931: by the kinetic energy term in the action: if the action reads
932: ${\cal S} ={1\over 2} \int d\tau g_{IJ}\dot{x}^I \dot{x}^J + \dots$,
933: where $x^I$ are the coordinates, then $g_{IJ}$ is the metric on
934: the space of coordinates. The sum over $I$
935: includes integration
936: over $\sigma$ in our case.
937: This superspace metric is needed for quantizing
938: the theory, for example in the coordinate representation one
939: needs an inner product on Hilbert space and this
940: involves integration over coordinates.
941: The determinant of $g_{IJ}$ is needed in
942: order to define this integral (see Appendix 4).
943:
944: The simplest way to identify the physical degrees of freedom
945: is to choose a gauge, for example $A$=constant, $A_i=0$.
946: For bulk particles, we can then read off the metric
947: on coordinate space from the action (\ref{eq:parta}) -
948: in this case it is simply the background
949: metric itself.
950:
951: We have already derived the Hamiltonian for winding states, and
952: showed how through the use of Dirac brackets
953: the $\theta$ coordinate may be discarded. If we choose
954: the gauge
955: $A=1$, $A_i=0$
956: in the Hamiltonian (\ref{eq:totalh1}) with constraints
957: given in (\ref{eq:redcon}), we can
958: construct the
959: corresponding gauge-fixed action:
960: \be
961: {\cal S}_{gf}= \int d\tau d^{p-1}\sigma \, \,
962: {1\over 2} \left(\dot{x}^\mu \dot{x}^\mu \eta_{\mu \nu}
963: -\mu_{p-1}^2 \theta_0^2 t^2 {\rm Det}( {x}^\mu_{,i} x^\nu_{,j}
964: \eta_{\mu \nu})\right)
965: \labeq{actw}
966: \ee
967: where $\mu,\nu $ run over $0$ to $d-1$ and $i,j$ run from
968: $1$ to $p-1$. One may check that the classical equations
969: following from the action (\ref{eq:actw}) are the
970: correct Lagrangian equations for the $p$-brane in a certain
971: worldvolume coordinate system and that these equations preserve the
972: constraints (\ref{eq:redcon}) (see Appendix 3).
973:
974: The metric on the space of coordinates may be inferred from
975: the kinetic term in (\ref{eq:actw}), and it is just
976: the Minkowski metric.
977: In contrast, as discussed, the metric
978: on the space of coordinates for bulk states involves
979: the full background metric (\ref{eq:backg}) which
980: degenerates at $t=0$.
981: The difference means that whereas the
982: quantum fields describing winding states are regular in
983: the neighborhood of $t=0$, those describing bulk
984: states exhibit
985: logarithmic
986: divergences. In the penultimate section of this paper
987: we argue that these divergences are plausibly the origin of the
988: bad perturbative behavior displayed by strings
989: and particles propagating on Lorentzian orbifolds, behavior
990: we do not expect to be exhibited in M2-brane winding states
991: in M theory.
992:
993:
994: \section{Toy Model: Winding Strings in ${\cal M}_C \times R^{d-1}$}
995:
996: Before approaching the problem of quantizing winding membranes,
997: we start with a toy model
998: consisting of winding string states propagating in
999: ${\cal M}_C \times R^{d-1}$. This problem has also been
1000: considered by others~\cite{bachas,pioline} and
1001: in more detail than we shall
1002: do here. They point out and exploit interesting analogies with
1003: open strings in an electric field. Our focus will
1004: be somewhat different and will serve mainly as a
1005: warmup for case of winding M2-branes
1006: which we are
1007: more interested in.
1008:
1009: Strings winding uniformly
1010: around the compact $\theta$ dimension in (\ref{eq:backg}) appear as
1011: particles from the
1012: $d-$dimensional point of view. To study the classical
1013: behavior of these particles, it is convenient to
1014: start from the Nambu action for the string,
1015: \be
1016: {\cal S} = - \mu \int d^2\sigma \sqrt{-{\rm Det}(
1017: \partial_\alpha x^\mu \partial_\beta x^\nu g_{\mu \nu})},
1018: \labeq{string}
1019: \ee
1020: where $\mu$ is the string tension (to avoid clutter
1021: we set $\mu_1=\mu$ for
1022: the remainder of this section). The string worldsheet coordinates
1023: are $\sigma^\alpha=(\tau,\sigma)$.
1024:
1025: For the winding states we consider, we can set $\theta =\sigma$,
1026: so $0\leq \sigma\leq \theta_0$. We insist that the other space-time
1027: coordinates of the string $x^\mu=(t,\vec{x})$ do not depend on
1028: $\sigma$. It is convenient also to choose the gauge
1029: $t=\tau$, in which the action (\ref{eq:string}) reduces to
1030: \be
1031: {\cal S} = - \mu \theta_0 \int dt |t| \sqrt{1-\dot{\vec{x}}\,^2},
1032: \labeq{partnam}
1033: \ee
1034: in which $t$ is now the time, not a coordinate. This is
1035: the usual square root action for
1036: a relativistic particle, but with a time-dependent
1037: mass $\mu \theta_0 |t|$. The
1038: canonical momentum is $\vec{p}= \mu \theta_0
1039: |t| \dot{\vec{x}}/\sqrt{1-\dot{\vec{x}}\,^2}$ and the
1040: classical Hamiltonian generating evolution in the time $t$ is
1041: $H= \sqrt{\vec{p}\,^2+(\mu \theta_0 t)^2}$. This is
1042: regular at $t=0$, indicating that the classical equations
1043: should be regular there.
1044:
1045: Due to translation invariance, the canonical momentum
1046: $\vec{p}$
1047: is a constant of the motion. Using this, one obtains the
1048: general solution
1049: \be
1050: \vec{x}= \vec{x_0} + {\vec{p}\over \mu \theta_0}
1051: {\rm sinh}^{-1}\left(\mu \theta_0 t/|\vec{p}|\right),
1052: \qquad -\infty <t <\infty,
1053: \labeq{particle}
1054: \ee
1055: according to which the particle moves smoothly through
1056: the singularity. At early and late times the large mass
1057: slows the motion to a crawl.
1058: However, at $t=0$ the particle's mass disappears and
1059: it instantaneously
1060: reaches the speed of light.
1061: The key point for us is that these winding states
1062: have completely unambiguous evolution across $t=0$, even though
1063: the background metric (\ref{eq:backg}) is singular there.
1064:
1065: Now we turn to quantizing the theory, as a warmup for
1066: the membranes we shall consider in the next section.
1067: The relevant classical Hamiltonian was given
1068: in (\ref{eq:hemp}): it describes a point particle
1069: with a mass $\mu \theta_0 |t|$. In a general background
1070: space-time, ordering ambiguities appear, which are
1071: reviewed in Appendix 4. However, in the case at hand,
1072: there are no such ambiguities.
1073: The metric on the space of
1074: coordinates is the Minkowski metric $\eta_{\mu \nu}$.
1075: The standard expression for the momentum operator
1076: $p_\nu = -i \hbar (\partial /\partial x^\nu)$, and
1077: the
1078: Hamiltonian $H$ given in (\ref{eq:hemp})
1079: are clearly hermitian under
1080: integration over coordinate
1081: space, $\int d^d x$. Finally, the background
1082: curvature $R$ vanishes
1083: for our background so there
1084: is no curvature term ambiguity either.
1085:
1086: Quantization now proceeds by setting
1087: $p_\mu= - i \partial_\mu$ (we use units in which $\hbar$ is unity)
1088: in the Hamiltonian constraint (\ref{eq:hemp}) which
1089: is now an operator acting on the quantum field $\phi$.
1090: Fourier transforming with respect to
1091: $\vec{x}$, we obtain
1092: \be
1093: \ddot{\phi} = -\left(\vec{p}\,^2+ (\mu \theta_0 t)^2\right) \phi,
1094: \labeq{KG}
1095: \ee
1096: {\it i.e.,} the Klein Gordon equation for a particle with a mass
1097: $\mu \theta_0 |t|$.
1098:
1099: Equation (\ref{eq:KG}) is the parabolic cylinder equation. Its
1100: detailed properties are discussed in Ref.~\cite{AS},
1101: whose notation we follow.
1102: We write the time-dependent frequency as $\omega\equiv
1103: \sqrt{\vec{p}\,^2 +(\mu \theta_0 t)^2}$.
1104: At large times $\mu \theta_0 |t|\gg |\vec{p}\,|$,
1105: $\omega $ is slowly varying:
1106: $\dot{\omega}/\omega^2 \ll 1$ so all modes follow WKB
1107: evolution. The general solution behaves as a
1108: linear combination of
1109: $\omega^{-{1\over 2}}$exp$(\pm i \int \omega dt)
1110: \approx
1111: t^{-{1\over 2}} $exp$\left
1112: (\pm i {1\over 2}\left(\mu \theta_0 t^2
1113: +(\vec{p}\,^2{\rm ln}t)/(\mu \theta_0))\right)\right)$.
1114:
1115: For large momentum, $\vec{p}\,^2\gg \mu \theta_0$,
1116: the WKB approximation
1117: remains valid for all time since $\dot{\omega}/\omega^2$
1118: is never large. In the WKB approximation there is no
1119: mode mixing and no particle production.
1120: Therefore for large momentum one expects little
1121: particle production. Departures from
1122: WKB are nonperturbative in
1123: $\dot{\omega}/\omega^2$, as explicit calculation verifies,
1124: the result scaling as $\sim$
1125: exp$(-|\omega^2/\dot{\omega}|_{max})
1126: \sim $ exp$(-\vec{p}\,^2/\mu \theta_0)$, at large $\vec{p}^2$.
1127:
1128: The parabolic cylinder functions which behave as positive and
1129: negative frequency modes at large times
1130: are denoted
1131: $E(a,x)$ and $E^*(a,x)$,
1132: where $x=\sqrt{2\mu \theta_0} t$ and
1133: $a=-\vec{p}\,^2/(2 \mu \theta_0)$. For
1134: positive $x$ they behave respectively
1135: as $x^{-{1\over 2}}$ exp$(+ i({1\over 4}x^2-a {\rm ln} x)$ and
1136: $x^{-{1\over 2}}$ exp$(- i({1\over 4}x^2-a {\rm ln} x)$.
1137: Both $E(a,x)$ and $E^*(a,x)$ are analytic at $x=0$.
1138: They are
1139: uniquely continued to
1140: negative values through the relation
1141: \be
1142: E(a,-x)= -i e^{\pi a} E(a,x) +i \sqrt{1+e^{2\pi a}} E^*(a,x).
1143: \labeq{cont}
1144: \ee
1145: For $t<0$, $E(a,-\sqrt{2\mu \theta_0}t)$ is the positive frequency
1146: incoming mode. As we extend $t$ to positive values, (\ref{eq:cont})
1147: yields the outgoing solution consisting of a linear combination
1148: of the positive frequency solution $E^*$ and the negative
1149: frequency solution $E$. The Bogoliubov coefficient~\cite{birrell} $\beta$
1150: for modes of momentum $\vec{p}$ is read off from (\ref{eq:cont}):
1151: \be
1152: \beta=-i e^{\pi a} = -i e^{-{1\over 2} \pi \vec{p}\,^2/(\mu \theta_0)}.
1153: \labeq{beta}
1154: \ee
1155: The result is exponentially suppressed at large
1156: $\vec{p}$, hence, the
1157: total number of particles per unit volume
1158: created by passage through the singularity is
1159: \be
1160: \int {d^{d-1} \vec{p} \over (2\pi)^{d-1}} |\beta|^2
1161: = \left({\mu \theta_0 \over 2\pi}\right)^{d-1\over 2},
1162: \labeq{ppres}
1163: \ee
1164: which is finite and tends to zero as the rapidity of the
1165: brane collision is diminished.
1166:
1167: It is interesting to ask
1168: what happens if we attempt to attach the $t<0$ half of
1169: ${\cal M}_C \times R^{d-1}$, (\ref{eq:backg})
1170: with rapidity parameter $\theta_0^{in}$, to the upper half of
1171: ${\cal M}_C \times R^{d-1}$ with a different rapidity parameter
1172: ${\theta}_0^{out}$. After all,
1173: the field equations for general
1174: relativity break down at $t=0$ and, hence, there
1175: is insufficient Cauchy data to uniquely determine the solution
1176: to the future. Hence, it might seem that we have the freedom to
1177: attach a future compactified
1178: Milne with any parameter ${\theta}_0^{out}$, since this would
1179: still be locally flat away from $t=0$ and,
1180: hence, a legitimate string theory background. However
1181: it is quickly seen that this is not allowed.
1182: By matching the
1183: field $\phi$ and its first time derivative $\partial_t \phi$
1184: across $t=0$, we can determine the particle production
1185: in this case.
1186: We find that due to the jump in $\theta_0$,
1187: the Bogoliubov coefficient $\beta$ behaves like
1188: $((\theta_0^{in})^{1\over 2}-({\theta}_0^{out})^{1\over 2})/
1189: \sqrt{\theta_0^{in} \theta_0^{out}}$,
1190: at large momentum, independent of $\vec{p}$.
1191: This implies
1192: divergent particle production, and indicates that to lowest
1193: order one only obtains sensible results in the analytically-
1194: continued background, which has ${\theta}_0^{out}=\theta_0^{in}$.
1195: We conclude that to retain
1196: a physically sensible theory we must have ${\theta}_0^{out}=\theta_0^{in}$,
1197: at least at lowest order in the interactions.
1198:
1199: \section{Dynamics of Winding M2-branes}
1200:
1201: We now turn to the more complicated but far more interesting
1202: case of winding membranes in ${\cal M}_C \times R^{d-1}$.
1203: We have in mind eleven-dimensional M theory,
1204: where the eleventh, M theory dimension shrinks away to
1205: a point. When this dimension is small but static,
1206: well known arguments~\cite{witten}
1207: indicate that M theory should tend to a string theory:
1208: type IIA for circle compactification, heterotic
1209: string theory for orbifold compactification.
1210: It is precisely the winding membrane states we are
1211: considering which map onto the string theory states as
1212: the M theory dimension becomes small. What makes this
1213: case specially interesting is that the
1214: string theory states include the graviton and the dilaton.
1215: Hence, by describing string propagation
1216: across $t=0$ we are describing the propagation
1217: of perturbative gravity across a
1218: singularity, which as explained in Section II is
1219: an FRW cosmological singularity from the $d$ dimensional
1220: point of view.
1221:
1222: A winding 2-brane
1223: is a string from the $d$-dimensional point of view.
1224: As explained in section III, the Hamiltonian
1225: for such strings may be expressed in a general gauge
1226: as
1227: \be
1228: H=\int d\sigma \left({A\over 2} \left(\pi_\mu \pi_\nu \eta^{\mu \nu}
1229: +\mu_2^2 \theta_0^2 t^2 \eta_{\mu \nu} {x^\mu}' {x^\nu}'\right)
1230: + A^1 {x^\mu}' \pi_\mu \right),
1231: \labeq{totalh2}
1232: \ee
1233: where $A$ and $A^1$ are arbitrary functions of the
1234: worldsheet coordinates $\sigma$ and $\tau$, and prime
1235: represents the derivative with respect to $\sigma$.
1236: Here $\mu,\nu$ run over $0,1,..,d-1$, and primes denote derivatives
1237: with respect to $\sigma$.
1238: The Hamiltonian is supplemented by the following first class
1239: constraints
1240: \be
1241: \pi_\mu \pi_\nu \eta^{\mu \nu}
1242: +\mu_{2}^2 \theta_0^2 t^2 \eta_{\mu \nu} {x^\mu}' {x^\nu}'
1243: \approx 0;\qquad
1244: {x^\mu}' \pi_\mu\approx 0,
1245: \labeq{redconr}
1246: \ee
1247: which ensure the Hamiltonian is weakly zero. The latter
1248: constraint is the familiar requirement that the momentum density is normal to
1249: the string.
1250:
1251: The key point for us is that the Hamiltonian and the constraints
1252: are regular in the
1253: neighborhood of
1254: $t=0$, implying that for a generic class of worldsheet coordinates
1255: the solutions of the equations of motion are regular there.
1256:
1257: As we did with particles, it is instructive to examine the
1258: classical theory from the point of view of the Nambu
1259: action. We may directly infer the classical action for
1260: winding membranes on ${\cal M}_C \times R^{d-1}$ by
1261: setting
1262: $\theta=\sigma^2$ and $\partial_2 t=\partial_2
1263: \vec{x} =0$. The Nambu action for the 2-brane then
1264: becomes
1265: \be
1266: {\cal S} = - \mu_2\theta_0
1267: \int d^2\sigma |t| \sqrt{-{\rm Det}(
1268: \partial_\alpha x^\mu \partial_\beta x^\nu \eta_{\mu \nu})},
1269: \labeq{stringmem}
1270: \ee
1271: where $\sigma^\alpha=(\tau,\sigma)$ and $\mu$ runs over $0,..d-1$.
1272: This is precisely the action for a string
1273: (\ref{eq:string}) in a time-dependent
1274: background $g_{\mu \nu}= \theta_0 |t|\eta_{\mu \nu}$, the appropriate
1275: string-frame background corresponding to ${\cal M}_C \times R^{d-1}$
1276: in M theory~\cite{kosst}.
1277:
1278: As a prelude to quantization, let us discuss the classical
1279: evolution of winding membranes across $t=0$.
1280: We can pick
1281: worldsheet coordinates so that $x^0\equiv t=\tau$, and
1282: $\dot{\vec{x}}\cdot \vec{x}\,'=0$.
1283: In this gauge the Nambu action is:
1284: \be
1285: {\cal S} = - \mu_2 \theta_0 \int dt d\sigma |t| |\vec{x}\,'|
1286: \sqrt{1-\dot{\vec{x}}\,^2},
1287: \labeq{nambu}
1288: \ee
1289: and the classical equations are
1290: \be
1291: \partial_t \left(\epsilon \dot{\vec{x}}\right)
1292: = \partial_\sigma \left( {t^2\over \epsilon} \partial_\sigma \vec{x}\right),
1293: \qquad \partial_t \epsilon = t {(\vec{x}\,')^2\over \epsilon},
1294: \labeq{stringe}
1295: \ee
1296: where
1297: \be
1298: \epsilon = |t| \sqrt{\vec{x'}\,^2/(1-\dot{\vec{x}\,^2})}
1299: \labeq{stringf}
1300: \ee
1301: the energy density along the string is $\mu_2 \epsilon$.
1302: It may be checked that the solutions of these
1303: equations are regular at $t=0$, with the
1304: energy and momentum density being finite there.
1305: The local speed of
1306: the string hits unity at $t=0$, but
1307: there is no ambiguity in the resulting solutions.
1308: The data at $t=0$ consists
1309: of the string coordinates $\vec{x}(\sigma)$ and the
1310: momentum density $\vec{\pi}(\sigma) = \mu_2 \epsilon \dot{\vec{x}}$,
1311: which must be
1312: normal to $\vec{x}\,'(\sigma)$ but is
1313: otherwise arbitrary.
1314: This is the same amount of initial data as that
1315: pertaining at any other
1316: time.
1317:
1318: Equations (\ref{eq:stringf}) describe
1319: strings in
1320: a $d$ dimensional flat FRW cosmological
1321: background with $ds^2= a(t)^2 \eta_{\mu \nu}$ and
1322: scale factor $a(t) \propto t^{1\over 2}$.
1323: The usual cosmological
1324: intuition is helpful in understanding
1325: the string evolution. The comparison
1326: one must make is between the curvature scale on the
1327: string and the comoving Hubble radius $|t|$.
1328: When $|t|$ is larger than the comoving
1329: curvature scale on the string, the string oscillates as in flat
1330: space-time, with fixed proper amplitude and frequency.
1331: However, when $|t|$ falls below the string curvature scale,
1332: the string is `frozen' in comoving coordinates.
1333: This point of view is useful in understanding the
1334: qualitative behavior we shall discuss in the next section.
1335:
1336: There is one final point we wish to emphasize. In Section II we
1337: discussed the general Hamiltonian for a $p$-brane
1338: in curved space. As we have just seen, a winding M2-brane
1339: on ${\cal M}_C \times R^{d-1}$ has the same action as a string in
1340: the background $g_{\mu \nu}= |t|\eta_{\mu \nu}$. We could
1341: have considered this case directly using the methods
1342: of Section II. The constraints (\ref{eq:consta}) for this case
1343: read:
1344: \be
1345: (\theta_0 |t|)^{-1}\eta^{\mu \nu} \pi_\mu \pi_\nu + \mu_2^2 \theta_0 |t| \eta_{\mu
1346: \nu}
1347: {x^\mu}' {x^\nu}'\approx 0 ; \qquad {x^\mu}'\pi_\mu\approx 0,
1348: \labeq{strcon}
1349: \ee
1350: and the Hamiltonian would involve an arbitrary linear combination
1351: of the two. These constraints may be compared with those
1352: coming directly from our analysis of
1353: winding 2-branes, given by (\ref{eq:redcon})
1354: with $p=2$.
1355: Only the first constraint differs,
1356: and {\it only by multiplication by
1357: $\theta_0 |t|$}.
1358: For all nonzero $\theta_0 |t|$, the difference is insignificant:
1359: the constraints are equivalent.
1360: Multiplication of the Hamiltonian by any
1361: function of the canonical variables merely amounts to
1362: a re-definition
1363: of worldsheet time.
1364: This is just as it should be: dimensionally reduced
1365: membranes are strings. However, the membrane viewpoint is
1366: superior in one respect, namely that the background metric
1367: (\ref{eq:backg}) is non-singular in the physical
1368: coordinates (which do not include $\theta$).
1369: That is why
1370: the membrane Hamiltonian constraint is nonsingular for these states.
1371: The membrane viewpoint tells us we should
1372: multiply the string Hamiltonian
1373: by $\theta_0|t|$ in order to obtain a string theory which is
1374: regular at $t=0$.
1375: Without knowing about membranes, the naive reaction
1376: might have been to discard the string theory on the basis
1377: that the string frame metric is singular there.
1378:
1379:
1380: \section{Einstein Gravity versus an Expansion in $1/\alpha'$}
1381:
1382:
1383: We are interested in the behavior of M theory,
1384: considered as a theory of M2-branes, in the
1385: vicinity of $t=0$. The first question is, for what
1386: range of $|t|$ is the string description valid?
1387: The effective string tension $\mu_1$
1388: is given
1389: in terms of the M2-brane tension $\mu_2$ by
1390: \be
1391: \mu_1= \mu_2 \theta_0 |t|.
1392: \labeq{stringt}
1393: \ee
1394: The mass scale of stringy excitations is the string scale $\mu_1^{1\over 2}$.
1395: For the stringy description to be valid, this scale must
1396: be smaller than the mass of Kaluza-Klein excitations,
1397: $(\theta_0 |t|)^{-1}$. This condition reads
1398: \be
1399: |t|<\mu_2^{-{1\over 3}} \theta_0^{-1}.
1400: \labeq{stringtime}
1401: \ee
1402: The second question is: for what range of $|t|$ may
1403: the string theory be approximated by $d$-dimensional
1404: Einstein-dilaton gravity?
1405: %PAUL added sentence
1406: Recall that the string frame metric is $g_{\mu \nu} = a^2 \eta_{\mu \nu} = |t|
1407: \eta_{\mu \nu}$,
1408: so the curvature scale is $ \dot{a}/a \sim 1/|t|$.
1409: The approximation holds
1410: when the curvature scale is smaller than the string scale $\mu_1^{1\over 2}$,
1411: which implies
1412: \be
1413: |t|> \mu_2^{-{1\over 3}} \theta_0^{-{1\over 3}}.
1414: \labeq{stringtime2}
1415: \ee
1416: We conclude that for small $\theta_0$
1417: there are three relevant regimes.
1418: In units
1419: where the membrane tension $\mu_2$ is unity, they
1420: are as follows. For $|t| > \theta_0^{-1}$
1421: the description of $M$ theory in terms of
1422: eleven dimensional Einstein gravity should hold.
1423: As $|t|$ falls below $\theta_0^{-1}$,
1424: the size of the extra dimension falls below
1425: the membrane scale and we go over to the
1426: ten dimensional string theory description.
1427: In the cosmological scenarios of interest,
1428: the incoming state is very smooth~\cite{chaos,review},
1429: hence this state should be well-described by
1430: ten dimensional Einstein-dilaton gravity, which
1431: of course agrees with the eleven dimensional Einstein theory
1432: reduced in the Kaluza-Klein fashion.
1433: However, when $|t|$ falls below $ \theta_0^{-{1\over 3}}$,
1434: the Einstein-dilaton description fails and we must
1435: employ the fundamental description of strings in
1436: order to obtain regular behavior at $t=0$.
1437:
1438: Let us consider, then, the relevant string action.
1439: As we explained in Section IV, for winding M2-branes
1440: a good gauge choice
1441: is $A=1$, $A^1=0$.
1442: The corresponding gauge-fixed
1443: worldsheet action was given in (\ref{eq:actw}): for $p=2$ this reads
1444: \be
1445: {\cal S}_{gf}= \int d\tau d\sigma
1446: {1\over 2} \eta_{\mu \nu} \left(\dot{x}^\mu \dot{x}^\nu
1447: -\mu_2^2 \theta_0^2 t^2 {x^\mu}' {x^\nu}')\right).
1448: \labeq{newact}
1449: \ee
1450: where
1451: the fields $x^\mu = (t,\vec{x})$ depend on $\sigma$ and $\tau$.
1452: This action describes a two dimensional field theory
1453: with a quartic interaction.
1454:
1455: We confine ourselves to some preliminary remarks about
1456: the perturbative behavior of (\ref{eq:newact})
1457: before proceeding to
1458: non-perturbative calculations in the next section.
1459: In units where $\hbar$ is unity, the
1460: action must be dimensionless. From the $d$-dimensional point of view,
1461: the coordinates
1462: $x^\mu$ have dimensions of inverse mass and $\mu$ has dimensions
1463: of mass cubed, so $ \sigma/\tau$ must have dimensions of
1464: mass squared. However, the dimensional analysis relevant to
1465: the quantization of (\ref{eq:newact}) considered as a two dimensional field
1466: theory is quite different:
1467: the fields $x^\mu$ are dimensionless. From
1468: the quartic term we see that $\mu_2 \theta_0$ is a dimensionless coupling.
1469: suggesting that perturbation theory in
1470: $\mu_2 \theta_0$ should be
1471: renormalizable. The string tension at a time $t_1$ is
1472: $\mu_1 = \mu_2 \theta_0 |t_1|$; the usual $\alpha'$
1473: expansion is then an expansion in the Regge slope parameter $\alpha'
1474: = 1/(2 \pi \mu_1)$, i.e. in negative powers of $\mu_2 \theta_0$.
1475: Conversely, the perturbation expansion
1476: we are discussing
1477: is an {\it expansion in inverse powers of $\alpha'$}.
1478:
1479: These considerations point the way to the resolution of
1480: an apparent conflict between two facts. Winding M2-brane evolution
1481: is as we have seen smooth through $t=0$. We
1482: expect M2-branes to be described by strings near $t=0$.
1483: The low energy approximation to string theory
1484: is Einstein-dilaton gravity. Yet,
1485: as noted above, in Einstein-dilaton gravity,
1486: generic perturbations
1487: diverge logarithmically with $|t|$ as $t$ tends to zero.
1488: The resolution of the paradox
1489: is that general relativity is the leading term in
1490: an expansion in $\alpha'$ for the string theory.
1491: As we have shown, however, the
1492: good behavior of string theory is only apparent
1493: in the opposite expansion, in {\it inverse} powers
1494: of $\alpha'$.
1495:
1496: In order to evolve
1497: the incoming
1498: state defined at large $|t|$ where
1499: general relativity
1500: is a good
1501: description, through small $|t|$ where string theory is still
1502: valid but general relativity fails,
1503: we must match the standard $\alpha'$
1504: expansion onto the new $1/\alpha'$ expansion we have been
1505: discussing. We defer a detailed discussion of
1506: this fascinating issue to a future publication.
1507:
1508: The action (\ref{eq:newact}) describes a string with a time-dependent
1509: tension, which goes to zero at $t=0$. There is an extensive literature
1510: on the zero-tension limit of string theory (see for example \cite{devega}),
1511: in Minkowski space-time. At zero tension the action
1512: is (\ref{eq:newact}) but without the second term. This is the action
1513: for an infinite number of massless particles with no interactions.
1514: Quantum mechanically, there is no central charge and no critical
1515: dimension. However, as the tension is introduced, the usual
1516: central charge and critical dimension appear\cite{nicolaidis}.
1517:
1518:
1519: \section {Worldsheet Instanton Calculation of Loop Production}
1520:
1521: The string theory we are discussing, with action (\ref{eq:newact}),
1522: is nonlinear and therefore difficult to solve. We can still
1523: make substantial progress on questions of physical interest by
1524: employing nonperturbative instanton methods.
1525: One of the most interesting questions
1526: is whether one can calculate the quantum production of
1527: M2-branes as the universe passes through the big crunch/big bang
1528: transition. As we shall now show, this is indeed possible
1529: through Euclidean instanton techniques.
1530:
1531: %In order to make progress, we now restrict attention to
1532: %the set of winding membrane/string states with axial symmetry {\it i.e.,}
1533: %circular loops, with a general centre of mass momentum
1534: %$\vec{p}_{cm}$. In the $t=\tau$, $\dot{\vec{x}}.\vec{x}'=0$ gauge,
1535: %the gauge conditions require that the plane of
1536: %the loop be perpendicular to $\vec{p}_{cm}$.
1537: %Note that even this limited class of string states is
1538: %of substantial interest, since in flat space-time
1539: %it includes the string
1540: %theory dilaton.
1541: %For a string loop of radius $R$
1542: %the Hamiltonian constraint is calculated from (\ref{eq:memcon})
1543: %to be
1544: %\be
1545: %\vec{p}_{cm}^2 + P_R^2 + (\mu \theta_0 t R)^2 =0.
1546: %\labeq{loopham}
1547: %\ee
1548: %Covariantly quantizing as before, we
1549: %obtain the field equation,
1550: %\be
1551: %\ddot{\phi} = -\left(-\vec{p}^2 -\partial_R^2 + (\mu \theta_0 R t)^2 \right)\phi
1552: %\labeq{circeq}
1553: %\ee
1554: %where $\nabla$ is the gradient with respect to the loop's
1555: %centre of mass
1556: %coordinate.
1557: %The second two
1558: %terms on the right hand side comprise the
1559: %Hamiltonian for the internal oscillation of the loop:
1560: %they describe
1561: %a harmonic oscillator with a time-dependent
1562: %frequency $\omega \propto |t|$. As before, we shall
1563: %temporarily adopt units in which $\mu \theta_0=1$ and then restore
1564: %$\mu \theta_0$ by dimensions in the final results.
1565: %
1566: %The solutions of (\ref{eq:circeq}) have simple behavior at
1567: %at large times.
1568: %The harmonic oscillator Hamiltonian
1569: %has eigenmodes which are functions of
1570: %$R |t|^{1\over 2} $ ({\it i.e.,} proper radius in string frame),
1571: %and eigenvalues $(2n+1) |t|$ for $n=0,1,\dots$.
1572: %These eigenvalues are large and
1573: %slowly varying at large $|t|$, so the adiabatic approximation holds.
1574: %The solutions
1575: %of (\ref{eq:circeq}) tend to the separable form
1576: %\be
1577: %\phi(\vec{x},R,t) \approx \sum_{\vec{p} n=0,1,\dots}
1578: %a_{n,\vec{p}} \,e^{i \vec{p}.\vec{x}}\, \psi_n(r)\, \chi_{n,p}(t)\quad +\quad{\rm h.c}
1579: %\labeq{gensol}
1580: %\ee
1581: %where $\vec{x}$ is the centre of mass coordinate of
1582: %the loop, with corresponding momentum $\vec{p}$,
1583: %$\psi_n(r)$ are harmonic oscillator eigenfunctions,
1584: %$\chi_{n,p}(t)$ is a properly
1585: %normalized positive frequency solution of
1586: %the ordinary differential equation,
1587: %\be
1588: %\ddot{\chi} = -\left(\vec{p}^2 + (2n+1) |t| \right) \chi
1589: %\labeq{chieq}
1590: %\ee
1591: %and the $a_n$ are the annihilation operators for the corresponding
1592: %excited loop states.
1593: %By
1594: %changing variables to $r=R |t|^{1\over 2}$ and
1595: %$T= |t|^{3\over 2}$, one may verify that
1596: %the approximation involved in (\ref{eq:gensol}) requires
1597: %only that $t >>1$.
1598:
1599: %Calculating the production of loop states
1600: %induced by the passage through $t=0$ requires
1601: %solving equation (\ref{eq:circeq}) with initial conditions
1602: %consisting at large negative times of a sum of
1603: %normalized harmonic oscillator states multiplied by the
1604: %appropriate positive frequency mode, for each value
1605: %of the incoming momentum $\vec{p}$. We then solve
1606: %equation (\ref{eq:circeq}) across $t=0$ to large positive
1607: %times and re-express the solution in terms of the
1608: %positive and negative frequency modes pertaining for
1609: %large positive times. The number of excitations
1610: %found in each harmonic oscillator state, and at each
1611: %momentum $\vec{p}$ is given by the square of the
1612: %Bogoliubov coefficient $\beta$ being the coefficient
1613: %of the outgoing negative frequency mode.
1614: %It is in principle straightforward to calculate $\beta$
1615: %numerically, but we shall not do so here.
1616:
1617: %Instead, we shall develop an approximate
1618: %analytic method yielding the
1619: %Bogoliubov coefficient from an imaginary time
1620: %classical solution {\it i.e.,} an
1621: %instanton.
1622:
1623:
1624: First, let us reproduce the result obtained in our toy model
1625: of winding string production, equation
1626: (\ref{eq:beta}).
1627: Equation (\ref{eq:KG}) may be
1628: re-interpreted
1629: as a time-independent Schrodinger equation, with $t$ being
1630: the coordinate, describing an
1631: over-the-barrier wave in a upside-down harmonic potential. The
1632: Bogoliubov
1633: coefficient is then just the ratio of the
1634: reflection coefficient $R$ to the transmission coefficient $T$.
1635: For large momentum
1636: $|\vec{p}\,|$, $R$ is exponentially small and $T$ is close
1637: to unity
1638: since the WKB
1639: approximation holds.
1640: To compute $R$ we employ the following
1641: approach, described in the
1642: book by Heading~\cite{heading}. (For related
1643: approximation schemes, applied to string production,
1644: see Refs.~\cite{martinec, gubser}.)
1645:
1646: \begin{figure}
1647: {\par\centering
1648: \resizebox*{3.in}{3.in}{\includegraphics{contour.eps}} \vskip .1in \par}
1649: \caption{The contour for computation of Bogoliubov coefficients
1650: in string/membrane production.
1651: }
1652: \labfig{contour}
1653: \end{figure}
1654: \noindent
1655:
1656:
1657:
1658: The method is to analytically
1659: continue the WKB approximate solutions in the complex
1660: $t$-plane. Defining the WKB frequency,
1661: %NNN introduce mu theta0
1662: $\omega=\sqrt{p^2+(\mu \theta_0 t)^2}$, one
1663: observes there are zeros at $t=\pm i p/(\mu \theta_0)$, where the
1664: WKB approximation must fail and where the WKB approximate
1665: solutions possess branch cuts. Heading shows that
1666: one can compute the reflection coefficient by
1667: running the branch cut from $t=-ip/(\mu \theta_0)$
1668: up the imaginary $t$ axis and out along the negative real $t$ axis.
1669: Then if one continues an incoming
1670: WKB solution, defined below the cut, in from $t=-\infty$,
1671: below the
1672: branch point at $t=-ip/(\mu \theta_0)$ and back out towards
1673: $t=-\infty$ just above the cut, it becomes the outgoing reflected
1674: wave. In the leading WKB approximation, the wave is
1675: given by $w^{-{1\over 2}} e^{-i \int w dt}$. Continuing this
1676: expression along the stated contour, shown in Figure 1,
1677: the exponent acquires a real
1678: contribution on the parts close to the
1679: imaginary axis, where $t=-i\tau$. Hence, the magnitude
1680: of the reflection amplitude $R$ is given in the first approximation
1681: by $e^{(-2 \int_0^p w d \tau)}$ where the factor
1682: of two arises from the two contributions on either side of the
1683: axis.
1684: With $w=\sqrt{p^2-(\mu \theta_0 \tau)^2}$, one easily sees that
1685: the exponent agrees precisely with that in (\ref{eq:beta}).
1686:
1687: A more direct method for getting the exponent in the
1688: Bogoliubov coefficient is to start
1689: not from the field equation (\ref{eq:KG}) but from the
1690: original action for the fundamental string, (\ref{eq:string}).
1691: We look for an imaginary-time solution ({\it i.e.,} an
1692: instanton) corresponding to the WKB continuation described in the
1693: previous section. The
1694: gauge-fixed particle Hamiltonian was given in (\ref{eq:hemp}), as
1695: $H= {1\over 2} \left(\vec{p}\,^2-p_0^2 +(\mu \theta_0 t)^2\right)$.
1696: The corresponding gauge-fixed action is, in first order form,
1697: \be
1698: {\cal S} = \int d \tau (- p_0 \dot{t} + \vec{p} \cdot \dot{\vec{x}} -H),
1699: \labeq{stract}
1700: \ee
1701: where as usual dots over a variable denote $\tau$ derivatives.
1702: Before continuing to imaginary time, it is important
1703: to realize that the spatial momentum $\vec{p}= \dot{\vec{x}}$
1704: is conserved (by translation invariance). Hence, all states,
1705: and in particular the asymptotic states we want, are
1706: labeled by $\vec{p}$. We are
1707: are interested in the transition
1708: amplitude for fixed initial and final $\vec{p}$,
1709: not $\vec{x}$, and we must use the appropriate action
1710: which is not (\ref{eq:stract}), but rather
1711: \be
1712: {\cal S} = \int d \tau (- p_0 \dot{t} - \dot{\vec{p}} \cdot \vec{x} -H),
1713: \labeq{stractm}
1714: \ee
1715: related by an integration by parts. The
1716: $\dot{\vec{p}}\cdot \vec{x}$ term contributes only a phase in the
1717: Euclidean path integral (because $\vec{p}$ remains
1718: real) and the $\vec{x}$ integration produces a delta
1719: function for
1720: overall momentum conservation. Notice that if we
1721: instead
1722: had used the naive action $\int d\tau {1\over 2}
1723: \dot{\vec{x}}\,^2$, we would have obtained a $\vec{p}\,^2$ term in
1724: the Euclidean action of
1725: the opposite sign. Similar considerations have been
1726: noted elsewhere~\cite{coleman}.
1727:
1728: Now we continue the action (\ref{eq:stractm})
1729: to imaginary time, setting $t=-it_E$
1730: and $\tau=-i\tau_E$. Eliminating $p_0$, the Euclidean action
1731: ${\cal S}_E \equiv -i{\cal S}$) is found to be
1732: \be
1733: {\cal S}_E= \int d \tau_E \left({1\over 2}
1734: \left( \dot{t}^2_E -(\mu \theta_0 t_E)^2 +\vec{p}\,^2\right)
1735: + i \dot{\vec{p}}\cdot\vec{x}\right).
1736: \labeq{eucac}
1737: \ee
1738: where dots now denote derivatives with respect to $\tau_E$.
1739: The amplitude we want involves $t_E$ running from $0$
1740: to $|\vec{p}\,|$ and back again: (\ref{eq:eucac}) is just the
1741: action for a simple harmonic oscillator and the required
1742: instanton
1743: is $t_E= p \,{\rm cos} (\mu \theta_0 \tau_E)$,
1744: $-\pi/2 <\mu \theta_0 \tau_E < \pi/2$.
1745: The corresponding Euclidean action is
1746: \be
1747: {\cal S}_E= {\pi\over 2} {\vec{p}\,^2\over \mu \theta_0},
1748: \labeq{eucacr}
1749: \ee
1750: giving precisely the exponent in (\ref{eq:beta}).
1751:
1752: Now we wish to apply this method to calculating the production
1753: of winding membrane states,
1754: described by the action
1755: (\ref{eq:newact}). As in the case of
1756: analogous calculations of vacuum bubble nucleation
1757: within field theory \cite{bubbles},
1758: it is plausible that objects with the greatest
1759: symmetry are produced since non-symmetrical deformations
1760: will generally yield a larger Euclidean action. Therefore
1761: one might guess that the
1762: dominant production mechanism is the production of
1763: circular loops. Let us start by considering this case.
1764: The constraint $\pi_\mu {x^\mu}'=0$ implies
1765: that the plane of such loops must be perpendicular to their
1766: center of mass momentum $p_\mu$. As in the particle production
1767: process previously considered, loops must be produced in
1768: pairs carrying equal and opposite momentum. The Hamiltonian
1769: for such circular loops is straightforwardly found to be
1770: $H= {1\over 2} \left(\vec{p}\,^2 +
1771: p_R^2 -p_0^2 +(2 \pi R \mu \theta_0 t)^2\right)$,
1772: Following
1773: the same steps that led to (\ref{eq:eucac}),
1774: we infer that the appropriate Euclidean action is
1775: \be
1776: {\cal S}_E= \int d \tau_E \left({1\over 2}
1777: \left( \dot{t}^2_E + \dot{R}^2 -(2 \pi \mu_2 \theta_0 R t_E)^2 +\vec{p}\,^2\right)
1778: + i \dot{\vec{p}}\cdot\vec{x}\right).
1779: \labeq{eucacmem}
1780: \ee
1781: This action describes two degrees of freedom $t_E$ and $R$ interacting
1782: via a positive potential $t_E^2 R^2$. Up to the trivial symmetries
1783: $t_E\rightarrow -t_E$, $R\rightarrow -R$, there is only one
1784: classical solution which satisfying the boundary conditions we
1785: want, namely starting and ending at $t_E=0$, and running up to
1786: the zero of the WKB frequency function at $2 \pi
1787: \mu \theta_0 R t_E = |\vec{p}\,|$. This solution has $T_E=R$ and
1788: the Euclidean action is found to be
1789: \be
1790: {\cal S}_E= {(2 |\vec{p}\,|)^{3\over 2}\over (2 \pi \mu_2 \theta_0)^{1\over 2} }
1791: \int_0^1 dx \sqrt{1-x^4}.
1792: \labeq{eucacev}
1793: \ee
1794: where the last integral is $
1795: \Gamma[{1\over 4}]^2/(6\sqrt{2 \pi})$, a constant of order unity
1796: which we shall denote $I$.
1797:
1798: The Euclidean action
1799: grows like $|\vec{p}\,|^{3\over 2}$ at large momentum:
1800: this means that the total production of loops is finite.
1801: Neglecting a possible numerical pre-factor
1802: in the Bogoliubov coefficient,
1803: we can estimate the number density
1804: of loops produced per unit volume,
1805: \be
1806: n \sim
1807: \int {d^{d-1} \vec{p} \over (2\pi)^{d-1}}
1808: e^{-2 S_E}
1809: = (\mu_2 \theta_0)^{(d-1)/3} {2^{13-7d\over 3} \Gamma\left(2(d-1)/3\right)
1810: \over 3 \pi^{d-1\over 6}
1811: I^{2(d-1)/3} \Gamma\left((d-1)/2\right)}
1812: \labeq{loops}
1813: \ee
1814: where $I$ is given above.
1815:
1816: From the
1817: instanton solution, the
1818: characteristic size of the loops and the time when they are produced
1819: are both of the same order,
1820: $R \sim |t| \sim (|\vec{p}\,|/\mu_2 \theta_0)^{1\over 2}\sim
1821: (\mu_2 \theta_0)^{-1/3}$.
1822: The effective string tension
1823: when they are produced is $\sim \mu_2 \theta_0 |t|
1824: \sim (\mu_2 \theta_0)^{2\over 3}$.
1825:
1826: We have restricted attention so far to the production of
1827: circular loops. It is also
1828: important to ask whether long, irregular strings are also
1829: copiously produced. Even though such strings would be
1830: disfavoured energetically, there is
1831: an exponentially large density of available states which
1832: could in principle compensate.
1833: An estimate may be made along the lines of
1834: Ref.~\cite{gubser}, by simply replacing
1835: $\vec{p}\,^2$ with $
1836: \vec{p}\,^2 +\mu_1 N$, where $\mu_1$ is the effective
1837: string tension and $N$ is the level number
1838: of the string excitations.
1839: This picture only makes sense for times greater than
1840: the string time, so we use the tension at
1841: the string time, $\mu_1\sim
1842: (\mu_2 \theta_0)^{2/3}$.
1843: The density of string states
1844: scales as $e^{\sqrt{N}}$
1845: hence one should replace (\ref{eq:loops})
1846: with a sum over $N$:
1847: \be
1848: \sim \sum_N e^{\sqrt{N}}
1849: \int d^{d-1} \vec{p} \,
1850: e^{-\left((\vec{p}\,^2+\mu_1 N)^{3\over 4}/\mu_1^{3\over 4}\right)}.
1851: \labeq{long}
1852: \ee
1853: The $N^{3\over 4}$ beats the $\sqrt{N}$ so the sum is
1854: dominated by modest $N$, indicating that the production
1855: of long strings is suppressed. According to this result,
1856: the universe emerges at the string time
1857: with of order one string-scale loop per string-scale
1858: volume, i.e. at a density comparable to but below
1859: the Hagedorn density.
1860:
1861:
1862: Another key question is whether
1863: gravitational back-reaction
1864: effects are likely to be significant
1865: at the transition. As the universe fills
1866: with string loops, what is their effect on the background geometry?
1867: We estimate this as follows. Consider a string loop
1868: of radius $R$ in M theory frame. Its mass $M$ is $2 \pi R$ times the
1869: effective string tension $\mu_2 L$, where $L$ is the
1870: size of the extra dimension. The effective Einstein-frame
1871: gravitational coupling (the inverse of the coefficient of
1872: $R/2$ in the Lagrangian density)
1873: is given
1874: by $\kappa_{d}^2 = \kappa_{d+1}^2/L$.
1875: The gravitational potential produced by such a loop
1876: in $d$ spacetime dimensions is\cite{maeda}
1877: \be
1878: \Phi = - \kappa_{d}^2{ M\over ((d-2)A_{d-2} R^{d-3}}
1879: \labeq{gravloop}
1880: \ee
1881: where $A_D$ is the area of the unit $D$-sphere,
1882: $A_D= 2 \pi{D+1\over 2}/\Gamma\left((D+1)/2\right)$.
1883: Specializing to the case of interest, namely
1884: 2-branes in eleven-dimensional M theory,
1885: the tension $\mu_2$ is related to the eleven dimensional
1886: gravitational coupling by a quantization condition relating
1887: to the four-form flux, reading\cite{duff}
1888: \be
1889: \mu_2^3= 2 \pi^2 /(n \kappa_{11}^2)
1890: \labeq{membten}
1891: \ee
1892: with $n$ an integer.
1893: Equations (\ref{eq:gravloop}) and (\ref{eq:membten})
1894: then imply that the typical gravitational potential around a
1895: string loop is
1896: \be
1897: \Phi = - {105\over 64 \pi \mu_2^2 R^6 n}
1898: \sim -(\mu_2^2 R^6 n )^{-1} \sim - \theta_0^2/n
1899: \labeq{deduce}
1900: \ee
1901: up to numerical factors.
1902:
1903: We conclude that the gravitational potential on the scale
1904: of the loops
1905: is of order $\theta_0^2$ and therefore is consistently small
1906: for small collision rapidity.
1907: Since the mean separation of the loops when they are
1908: produced is of order their size
1909: $R$, this potential $\Phi$ is
1910: the typical gravitational potential throughout space. Multiplying
1911: the $tt$ component of the background metric (\ref{eq:backg}) by
1912: $1+ 2\Phi$ and redefining $t$, we conclude that the outgoing metric
1913: has an expansion rapidity of
1914: order $\sim \theta_0(
1915: 1+ C \theta_0^2)$ with $C$ a constant of order unity. We conclude that
1916: for small
1917: $\theta_0$ the gravitational back-reaction due to string loop
1918: production is small. Note that loop production is
1919: a quantum mechanical effect taking place smoothly over
1920: a time scale of
1921: order $(\mu_2 \theta_0)^{-{1\over 3}}$. Therefore if the
1922: rapidity of the outgoing branes alters as we have estimated,
1923: it happens smoothly and not
1924: like the jump in $\theta_0$
1925: discussed at the end of section VI. Therefore the picture
1926: of loop production is consistent with the comments made there.
1927:
1928: \section{Strings on Lorentzian Orbifolds are not regular}
1929:
1930: We have shown that strings
1931: constructed as winding M2-branes
1932: on ${\cal M}_C\times R^{9}$
1933: are analytic in the neighborhood of
1934: $t=0$. This was the setup originally
1935: envisaged in the ekpyrotic model, where collapse of the
1936: M theory dimension was considered. Subsequently,
1937: a number of authors investigated the simpler
1938: case of string theory on ${\cal M}_C\times R^{8}$,
1939: considered as a Lorentzian orbifold solution
1940: of ten dimensional string theory.
1941: This is a simpler, but
1942: different setting, hence we expressed
1943: misgivings\cite{misgive,tolleyperts}
1944: about
1945: drawing conclusions
1946: from these reduced models.
1947: We shall now explain why the behavior in the
1948: Lorentzian orbifold models
1949: is significantly worse than in M theory
1950: and, hence, why no negative
1951: conclusion should be drawn on the basis of the
1952: failed perturbative calculations.
1953:
1954: Consider string theory on the background (\ref{eq:backg}).
1955: Let us choose the gauge $x^0=t=\tau$, and $g_{\mu \nu}
1956: \dot{x}^\mu \dot{x}'^\mu=0$. In this gauge we
1957: express the string solution as $\theta(t,\sigma)$ and $\vec{x}(t,\sigma)$.
1958: If a classical solution in this gauge possesses
1959: a singularity at $t=0$ in the complex $t$-plane, for all $\sigma$,
1960: then there
1961: can be no choice of worldsheet coordinates $\tau$ and
1962: $\sigma$ which can render the solution analytic in the neighborhood
1963: of $t=0$. For if such
1964: a choice existed,
1965: one could re-express $\tau$ in terms of $t$ and, hence,
1966: $\vec{x}(t,\sigma)$ would be analytic. We shall show that generically,
1967: for strings on Lorentzian orbifolds, the solutions possess
1968: logarithmic singularities {\it i.e.,}
1969: branch points, rendering them ambiguous as
1970: one circumvents the singularity in the complex $t$-plane.
1971:
1972: The demonstration is straightforward, and our argument
1973: is similar to that in earlier papers\cite{steif}. We are only
1974: interested in the classical equations of motion and
1975: we may compute the relevant Hamiltonian from
1976: the Nambu action,
1977: \be
1978: H\equiv \int d\sigma {\cal H}
1979: = \int d\sigma \sqrt{{\pi_\theta^2\over \mu t^2} +
1980: {\vec{\pi}^2 \over \mu} + \mu\left( (t {\theta'})^2+{\vec{x}'}\,^2\right)}.
1981: \labeq{loham}
1982: \ee
1983: The Hamiltonian equations allow generic solutions in which
1984: $\pi_{\theta}(\sigma)$ tends to any
1985: function of $\sigma$ as $t$ tends to zero. From its
1986: definition, the Hamiltonian density
1987: ${\cal H}$ then diverges as $t^{-1}$. The equation of motion for $\theta$
1988: is
1989: $\dot{\theta} = \pi_\theta/(t^2 {\cal H})$, implying
1990: that $\dot{\theta} \rightarrow \pm t^{-1}$ independent
1991: of $\sigma$. This implies a leading term
1992: $\theta \sim \pm {\rm log} t$,
1993: independent of $\sigma$.
1994: Recalling that ${\cal M}_C$ may be rewritten in
1995: flat coordinates by setting $T=t \cosh \theta$ and $Y=t \sinh \theta$,
1996: one readily understands this behavior. A geodesic in the
1997: $(T,Y)$ coordinates is just $Y= V T$, with $V$ a constant.
1998: At small $t$ this requires $e^\theta$ or $e^{-\theta}$ to
1999: diverge as $t^{-1}$, which is just the result we found.
2000:
2001: We conclude that generic solutions to the Hamiltonian
2002: equations possess branch points at $t=0$
2003: meaning that the solutions to the classical
2004: equations are ambiguous as one continues around $t=0$
2005: in the complex $t$-plane. This is a much
2006: worse situation that encountered for winding
2007: branes in M theory.
2008:
2009: The second problem occurs when we quantize and construct
2010: the associated field theory. Then the bad behavior at $t=0$
2011: corresponds to a diverging energy density which renders
2012: perturbation theory invalid.
2013: Since the previous problem
2014: can be seen even in pointlike states,
2015: let us focus attention
2016: on those. As we discuss in detail in Appendix 4, one needs
2017: to use the metric on the space of coordinates in order to construct the
2018: quantum Hamiltonian. In this case, the metric on the space of coordinates
2019: is the background metric, (\ref{eq:backg}). The field equation
2020: for point particles is then given from (\ref{eq:orderh}): Fourier
2021: transforming with respect to $\vec{x}$ and $\theta$, it reads
2022: \be
2023: \ddot{\phi} + {1\over t} \dot{\phi} = -\vec{k}\,^2 \phi - {k_\theta^2\over
2024: t^{2}} \phi,
2025: \labeq{eoff}
2026: \ee
2027: where $k_\theta$ is quantized in the usual way. This is
2028: the equation studied in earlier work\cite{tolley1} on
2029: quantum field theory on ${\cal M}_C\times R^{d-1}$. The generic
2030: solutions of (\ref{eq:eoff}) behave for small $t$
2031: as log~$t$ for $k_\theta=0$,
2032: or $t^{ik_\theta}$ for $k_\theta \neq 0$. In both cases, the
2033: kinetic energy density $\dot{\phi}^2$
2034: diverges as $t^{-2}$. Similar behavior is found
2035: for linearized vector and tensor fields on ${\cal M}_C\times R^{d-1}$.
2036: These divergences lead to the breakdown of perturbation theory
2037: in classical perturbative gravity,
2038: an effect which is plausibly the
2039: root cause of the
2040: bad behavior of the associated string theory scattering
2041: amplitudes~\cite{seibergetal}.
2042: As we have stressed, in the sector of M theory considered as
2043: a theory of membranes, describing perturbative gravity,
2044: this effect does not occur. The field
2045: equations are regular in the neighbourhood of $t=0$ and
2046: there is no associated divergence in the energy density.
2047:
2048: With hindsight, one can now see
2049: that directly constructing string theory on Lorentzian
2050: orbifolds sheds little
2051: light on the M theory case of interest in
2052: the ekpyrotic and cyclic models. Whilst the orbifolding
2053: construction provides a global map between incoming
2054: and outgoing free fields, it does not
2055: avoid the blueshifting effect which such fields
2056: generically suffer as they approach $t=0$, which
2057: seems to lead to singular behavior in the interactions
2058: (although Ref. \cite{ccosta} argues that a resummation may
2059: cure this problem).
2060:
2061: An alternate approach involving analytic continuation
2062: around $t=0$ has been simultaneously developed, but
2063: so far only implemented successfully in linearized
2064: cosmological perturbation theory\cite{tolley1,tolleyperts}.
2065: This method may in fact turn out to work even in the M theory
2066: context. The point
2067: is that by circumnavigating
2068: $t=0$ in the complex $t$-plane, maintaining a sufficient
2069: distance from the singularity, one may still retain
2070: the validity of linear perturbation theory and the
2071: use of the Einstein equations all the way along the
2072: complex time contour. The principles behind this would be
2073: similar
2074: to those familiar in the context of
2075: WKB matching via analytic continuation.
2076:
2077: In any case, the main point we wish to make is that now
2078: that we have what seems like a consistent
2079: microscopic theory for
2080: perturbative gravity, valid all the
2081: way through $t=0$, we have a reliable foundation
2082: for such investigations.
2083:
2084:
2085: \section{Conclusions and Comments}
2086:
2087:
2088: We have herein proposed an M theoretic model for the passage through a
2089: cosmological singularity in terms of a collision of orbifold planes
2090: in a compactified Milne ${\cal M}_C\times R^{9}$ background. The model
2091: begins with two empty, flat, parallel branes a string length apart approaching one
2092: another at constant rapidity $\theta_0$. With this initial condition, we have
2093: argued that the excitations naturally bifurcate into light winding M2-brane modes,
2094: and a set of massive modes including the Kaluza-Klein massive states.
2095: It is plausible that the
2096: massive modes decouple as their mass diverges.
2097: The light modes incorporate
2098: perturbative gravity and, hence, describe the space-time throughout the
2099: transition. Our finding that they are produced with finite density following
2100: dynamical equations that propagate smoothly through the
2101: transition supports the
2102: idea that this M-theory picture is
2103: well-behaved and predictive. Our model also
2104: suggests a string theory explanation
2105: of what goes wrong with
2106: Einstein gravity near the singularity: Einstein gravity is the leading
2107: approximation in an expansion in $\alpha'$, but the winding
2108: mode picture is a perturbative expansion in $1/\alpha'$.
2109:
2110:
2111: Our considerations have been almost entirely classical
2112: or semi-classical, although we believe the
2113: canonical approach we have adopted is a good starting point
2114: for a full quantum theory.
2115: Much remains to be done to fully
2116: establish the consistency of the picture we are advocating.
2117: In particular, we need to understand the sigma model
2118: in (\ref{eq:actnew}), and make sure that it is
2119: consistent quantum mechanically in the critical dimension.
2120: Second, we need to
2121: learn how to match the standard $\alpha'$ expansion to the
2122: $1/\alpha'$ expansion which we have argued
2123: should be smooth around $t=0$. Third, we need to incorporate
2124: string interactions. Although the vanishing of
2125: the string theory coupling around $t=0$ suggests that scattering
2126: plays a minor role,
2127: the task of fully constructing string perturbation
2128: theory in this background remains.
2129:
2130: Assuming these nontrivial tasks can be completed, can we
2131: say something about the significance of this example?
2132: The most obvious application is to the cyclic and ekpyrotic
2133: universe
2134: models which motivated these investigations,
2135: and which produce precisely the initial conditions required.
2136: As has been argued elsewhere~\cite{kosst,review}, the vicinity
2137: of the collision is well-modeled
2138: by compactified Milne times flat space.
2139: Within this model, the calculations reported here
2140: yield estimates of the reheat temperature immediately
2141: after the brane collision i.e. at the beginning
2142: of the hot big bang. Furthermore, if
2143: our arguments of Section VIII are correct, they
2144: should in principle provide
2145: complete matching rules for evolving cosmological
2146: perturbations
2147: through singularities of the type
2148: occurring in cyclic/ekpyrotic models. As we discussed
2149: briefly in Section X, it is plausible that this
2150: matching rule will recover the results obtained
2151: earlier for long-wavelength modes
2152: by the analytic continuation method~\cite{tolley1,tolleyperts},
2153: although that remains to be demonstrated in detail.
2154:
2155:
2156: Many other questions are raised by our work.
2157: Can we extend the treatment to other time-dependent
2158: singularities?
2159: The background considered here,
2160: ${\cal M}_C\times R^{9}$ is certainly very special
2161: being locally flat. Although the string frame metric
2162: is singular, it is conformally flat.
2163: One would like to
2164: study more generic string backgrounds corresponding to
2165: black holes, or
2166: Kasner/mixmaster
2167: spacetimes in general relativity.
2168: As we have argued, there is no reason to
2169: take the latter solutions seriously within M theory since they
2170: are only solutions of the low energy effective theory,
2171: which fails in the relevant regime. Nevertheless, they
2172: presumably have counterparts in M theory, and it remains a
2173: challenge to find them. We would like to believe that
2174: by constructing one consistent example, namely M theory on
2175: ${\cal M}_C\times R^{9}$ we would be opening the door
2176: to an attack on the generic case.
2177:
2178: Such a program is admittedly ambitious. However, should it
2179: succeed, we believe it
2180: would (and should) completely change our view of cosmology.
2181: If the best theories of gravity allow for
2182: a smooth passage through time-dependent
2183: singularities, this must profoundly
2184: alter our interpretation of the
2185: big bang, and of the major conceptual
2186: problems of the standard hot big bang cosmology.
2187:
2188:
2189: {\bf Acknowledgments}
2190: We would like to thank L. Boyle,
2191: S. Gratton, S. Gubser, G. Horowitz, S. Kachru, I. Klebanov,
2192: F. Quevedo, A. Tolley and P. Townsend for helpful remarks.
2193: This work was supported in part by PPARC, UK (NT and MP)
2194: and the US Department of Energy grants DE-FG02-91ER40671 (PJS).
2195: PJS is the Keck Distinguished Visiting
2196: Professor at the Institute for Advanced Study with support from
2197: the Wm.~Keck Foundation and the Monell Foundation and
2198: NT is the Darley Professorial Fellow at Cambridge.
2199:
2200: \bigskip
2201:
2202: \section*{Appendix 1: Canonical Treatment of $p$-Branes in Curved Space}
2203:
2204:
2205: In this Appendix we review the canonical treatment of
2206: $p$-Brane dynamics in curved space. A similar approach is
2207: taken in the recent work of Capovilla {\it et al.}~\cite{capovilla}.
2208: Earlier treatments include Refs.~\cite{townsend} and \cite{gutowski}.
2209:
2210: We start from the action in Polyakov form. No square roots appear and the
2211: ensuing general Hamiltonian involves only polynomial interactions.
2212: The action for a $p$-brane embedded in
2213: a background space-time with coordinates $x^\mu$ and metric
2214: $g_{\mu \nu}$ is:
2215: \be
2216: {\cal S} = -{1\over 2} \mu_p \int d^{p+1} \sigma
2217: \sqrt{-\gamma}\left(\gamma^{\alpha \beta}
2218: \partial_\alpha x^\mu \partial_\beta x^\nu g_{\mu \nu} -(p-1) \right),
2219: \labeq{pacta}
2220: \ee
2221: where $\mu_p$ is a mass per unit $p$-volume. The $p$-brane worldvolume
2222: has coordinates $\sigma^\alpha$, $\alpha=0,1,\dots,p$ in which the
2223: metric is $\gamma_{\alpha \beta}$.
2224: %The simplest case
2225: %is the action for a massive particle, $p=0$, with
2226: %$e^2\equiv -\gamma_{00}$ and
2227: %\be
2228: %{\cal S} = {1\over 2} m\int d \tau
2229: %\left(e^{-1}
2230: %\dot{x}^\mu \dot{x}^\nu g_{\mu \nu} -e\right),
2231: %\labeq{part}
2232: %\ee
2233: %Variation with respect to $e$ yields the constraint $e^2= -
2234: %\dot{x}^\mu \dot{x}^\nu g_{\mu \nu}$.
2235: %The canonical momentum is $p_\mu = m g_{\mu \nu}\dot{x}^\nu e^{-1}$ and
2236: %the constraint implies $g^{\mu \nu} p_\mu p_\nu = -m^2$.
2237:
2238:
2239: One way to proceed is to vary (\ref{eq:pacta})
2240: with respect to $\gamma_{\alpha \beta}$, hence, obtaining
2241: the constraint expressing the worldvolume metric
2242: $\gamma_{\alpha \beta}$ in terms of the
2243: the induced metric $\partial_\alpha x^\mu \partial_\beta x^\nu g_{\mu \nu}$.
2244: Substituting
2245: back into (\ref{eq:pacta}) one obtains the Nambu action for the
2246: embedding coordinates $x^\mu(\sigma)$ {\it i.e.,} $-\mu_p$ times the
2247: induced $p$-brane world volume.
2248:
2249: One can do better, however, by not eliminating the worldvolume
2250: metric so soon. Instead it is better to retain
2251: the $\gamma_{\alpha \beta}$ as independent variables and
2252: derive the corresponding
2253: Hamiltonian and constraints corresponding to evolution
2254: in worldvolume time $\tau$. It is convenient to
2255: express $\gamma_{\alpha \beta}$ in the form frequently
2256: used in canonical general
2257: relativity: the worldsheet line element is
2258: written
2259: \be
2260: \gamma_{\alpha \beta} d \sigma^\alpha d\sigma^\beta =
2261: (-\alpha^2 +\beta_k \beta^k) d\tau^2 + 2 \beta_i d\tau d\sigma^i
2262: +\overline{\gamma}_{ij} d\sigma^i d \sigma^j,
2263: \labeq{metric}
2264: \ee
2265: where
2266: $\sigma^i$, $i=1,\dots, p$ are the spatial worldvolume coordinates,
2267: $\beta^k$ is the shift vector and
2268: $\alpha$ is the lapse function. The good property of this
2269: representation is that the metric
2270: determinant simplifies: $\gamma=-\alpha^2 \overline{\gamma}$.
2271:
2272: In the following discussion we shall for the most part assume $p>1$ and
2273: then comment on the amendments needed for $p=0,1$.
2274: Using (\ref{eq:metric}) the action (\ref{eq:pacta}) becomes
2275: \be
2276: {\cal S} = -{1\over 2} \mu_p \int d\tau d^{p} \sigma \alpha
2277: \overline{\gamma}^{1\over 2}
2278: \left[ - {1\over \alpha^{2}} \dot{x}^\mu \dot{x}^\nu g_{\mu \nu}
2279: + 2 {\beta^i \over \alpha^2} \dot{x}^\mu x^\nu_{,i} g_{\mu \nu}
2280: + (\overline{\gamma}^{ij}-{\beta^i \beta^j \over \alpha^2} )
2281: {x}^\mu_{,i} x^\nu_{,j} g_{\mu \nu} -(p-1)\right].
2282: \labeq{actnew}
2283: \ee
2284: The canonical momenta conjugate to $x^\mu$ are found to be
2285: \be
2286: \pi_\mu = \mu_p {\overline{\gamma}^{1\over 2} \over \alpha}
2287: (\dot{x}^\nu - \beta^i x^\nu_{,i}) g_{\mu \nu}.
2288: \labeq{momeq}
2289: \ee
2290: No time derivatives of $\alpha$, $\beta^i$ and
2291: $\overline{\gamma}^{ij}$ appear in the Lagrangian, hence, the
2292: corresponding conjugate momenta vanish. In Dirac's language~\cite{dirac},
2293: these are the primary constraints.
2294: \be
2295: \pi_\alpha\approx 0, \qquad \pi_i\approx 0, \qquad\pi_{ij} \approx 0.
2296: \labeq{firstc}
2297: \ee
2298: Since Poisson brackets between momenta vanish, these constraints
2299: are first class.
2300: The total Hamiltonian $H$ then consists of the
2301: usual expression $H\equiv \int d^p \sigma \,[ \pi_A \dot{x}^A -L]$,
2302: \be
2303: H=\int d^p \sigma \left[{\alpha \over 2 \mu_p \overline{\gamma}^{1\over 2}}
2304: \pi_\mu \pi_\nu g^{\mu \nu} + \beta^i x^\mu_{,i} \pi_\mu +
2305: {\mu_p \alpha \overline{\gamma}^{1\over 2} \over 2}
2306: \left(\overline{\gamma}^{ij}{x}^\mu_{,i} x^\nu_{,j} g_{\mu \nu} -(p-1)\right)
2307: \right],
2308: \labeq{hama}
2309: \ee
2310: plus an arbitrary linear combination of the primary first class
2311: constraints (\ref{eq:firstc}).
2312:
2313: Additional secondary constraints are obtained from the Hamiltonian
2314: equations for $\dot{\pi}_\alpha$, $\dot{\pi}_i$ and $\dot{\pi}_{ij}$.
2315: Insisting these
2316: vanish as they must for consistency
2317: with (\ref{eq:firstc}), one finds
2318: \ba
2319: C&\equiv& \pi_\mu \pi_\nu g^{\mu \nu}
2320: +\mu_p^2
2321: \overline{\gamma}
2322: \left(\overline{\gamma}^{ij}{x}^\mu_{,i} x^\nu_{,j} g_{\mu \nu}
2323: -(p-1)\right)\approx 0,\cr
2324: C_i&\equiv&x^\mu_{,i} \pi_\mu\approx 0,\cr
2325: C_{ij}&\equiv& \gamma_{ij}-{x}^\mu_{,i} x^\nu_{,j} g_{\mu \nu} \approx 0.
2326: \labeq{secondc}
2327: \ea
2328: Following Dirac's procedure, we can now try to eliminate
2329: variables using second class constraints (constraints
2330: whose Poisson brackets with the other constraints
2331: does not vanish). In particular, it
2332: makes sense to eliminate $\gamma_{ij}$
2333: since the corresponding momenta $\pi_{ij}$ vanish weakly.
2334: It is easy to see that $C_{ij}$ are
2335: second class since their Poisson brackets with $\pi_{ij}$
2336: are nonzero.
2337: Hence, we eliminate
2338: $\gamma_{ij}$ using the $C_{ij}$ constraint, obtaining
2339: \be
2340: C\equiv \pi_\mu \pi_\nu g^{\mu \nu}
2341: +\mu_p^2 {\rm Det}( {x}^\mu_{,i} x^\nu_{,j} g_{\mu \nu})
2342: \approx 0, \qquad
2343: C_i\equiv x^\mu_{,i} \pi_\mu\approx 0,
2344: \labeq{secondcc}
2345: \ee
2346: as our new constraints, on the remaining variables
2347: $x^\mu$ and $\pi_\mu$. The matrix
2348: ${x}^\mu_{,i} x^\nu_{,j} g_{\mu \nu}$ is the
2349: induced spatial metric on the brane. When written this way, the
2350: $C$ and $C_i$ constraints
2351: have zero Poisson brackets with the remaining
2352: primary constraints $\pi_\alpha$ and $\pi_i$.
2353: A lengthy but straightforward calculation establishes
2354: that all Poisson brackets between
2355: $C$ and $C_i$ are weakly vanishing
2356: (see Appendix 2) and, hence, that we have a
2357: complete set of
2358: first class constraints consisting of $\pi_\alpha$,
2359: $\pi_i$, $C_i$ and $C$.
2360:
2361: The canonical Hamiltonian (\ref{eq:hama}) is now seen to be
2362: a linear combination of $C$ and $C_i$, with coefficients
2363: depending on $\alpha$ and $\beta^i$ respectively.
2364: The general Hamiltonian consists of a sum of this
2365: term plus an
2366: arbitrary linear combination of the first class
2367: constraints,
2368: $\int d \sigma (v_\alpha \pi_\alpha
2369: +v^i \pi_i)$ where $v_\alpha$ and $v_i$ are arbitrary
2370: functions of the worldvolume coordinates.
2371: From Hamilton's
2372: equations, one infers that $\dot{\alpha}=v_\alpha$,
2373: and $\dot{\beta}^i=v^i$. Therefore,
2374: $\alpha$ and $\beta$ are
2375: completely arbitrary functions of time. As Dirac
2376: emphasizes, one can then forget about $\alpha$,
2377: $\pi_\alpha$, $\beta^i$ and $\pi_i$ and just write the
2378: total Hamiltonian for the surviving coordinates as $x^\mu$ and
2379: $\pi_\mu$ as
2380: \be
2381: H=\int d^p\sigma \left({A\over 2}\left(\pi_\mu \pi_\nu g^{\mu \nu}
2382: +\mu_p^2
2383: {\rm Det}({x}^\mu_{,i} x^\nu_{,j} g_{\mu \nu})\right)+ A^i x^\mu_{,i} \pi_\mu
2384: \right),
2385: \labeq{totalh}
2386: \ee
2387: i.e. a linear combination of the constraints (\ref{eq:secondcc})
2388: with arbitrary coefficients $A$ and $A^i$.
2389: Different choices of $A$ and $A^i$ then correspond to different
2390: choices of worldvolume coordinates.
2391:
2392: We count the surviving physical
2393: degrees of
2394: freedom as follows. We start
2395: with the $2(d+1)$ coordinates $x^\mu$ and
2396: momenta $\pi^\mu$, each functions
2397: of the $p+1$ worldvolume coordinates. Then we impose
2398: the $p+1$ constraints
2399: $C=C_i=0$. Finally, in order to specify time evolution
2400: we must pick $p+1$
2401: arbitrary functions
2402: $A$ and $A_i$.
2403: The remaining physical degrees of
2404: freedom are $2(d+1)-2(p+1)=2(d-p)$ in number,
2405: the right number of transverse
2406: coordinates and momenta for the $p$-brane.
2407:
2408: We have tacitly assumed $p>1$ in the above analysis. The following
2409: minor amendments are needed for $p=0$ and $1$. For $p=0$,
2410: the worldvolume metric involves $\alpha$ only and one can
2411: ignore anything with an $i$ index except the determinant,
2412: which is replaced by unity. In particular there
2413: is no integration over $\sigma$ in the Hamiltonian, and
2414: the canonical momentum density $\pi_\mu$ is replaced by the
2415: momentum $p_\mu$.
2416: The only constraint is
2417: \be
2418: C\equiv p_\mu p_\nu g^{\mu \nu}+\mu_0^2\approx 0,
2419: \labeq{parti}
2420: \ee
2421: which is just the usual mass shell condition. The general
2422: Hamiltonian consists of an arbitrary function of $\tau$ times
2423: $C$.
2424:
2425: For $p=1$ the action (\ref{eq:pacta}) is invariant under conformal
2426: transformations of the worldvolume metric, and hence
2427: only two independent
2428: combinations of the three
2429: worldvolume metric variables appear in the decomposition
2430: (\ref{eq:metric}). The corresponding
2431: two momenta vanish and these are the primary constraints.
2432: Through Hamilton's equations one
2433: finds the
2434: following secondary constraints:
2435: \ba
2436: C&\equiv&\pi_\mu \pi_\nu g^{\mu \nu} +\mu_1^2 {{x}^\mu}' {x^\nu}' g_{\mu
2437: \nu}\approx 0,\cr
2438: C_1&\equiv&\pi_\mu {{x}^\mu}' \approx 0,
2439: \labeq{strincon}
2440: \ea
2441: where primes denote derivatives with respect to $\sigma^1\equiv \sigma$.
2442: The general Hamiltonian again
2443: takes the form (\ref{eq:totalh}), with $p=1$.
2444:
2445:
2446: \section*{Appendix 2: Poisson bracket algebra of the constraints}
2447:
2448: In the canonical theory \cite{dirac}, one considers arbitrary
2449: functions of the canonical
2450: variables, and of the time $\tau$.
2451: In our case, the canonical variables are fields
2452: $x^\mu(\sigma)$ and $\pi_\mu(\sigma)$
2453: depending upon $\sigma$, which is regarded
2454: as a continuous index labeling an infinite number
2455: of canonical variables. In particular, the constraints
2456: in (\ref{eq:secondcc}) are infinite in number. In
2457: this Appendix we shall show that the Poisson bracket
2458: algebra of the constraints closes, and, hence, that, in
2459: Dirac's terminology, they are first class.
2460:
2461: The Poisson bracket between any two quantities $M$ and $N$,
2462: which may be arbitrary functions of
2463: the canonical variables (local or nonlocal in $\sigma$)
2464: and of the time $\tau$, is
2465: given by
2466: \be
2467: \left\{M,N\right\}\equiv \int d^p\sigma \left(
2468: {\partial M \over \partial x^\mu(\sigma)}
2469: {\partial N\over \partial \pi_\mu(\sigma)}
2470: -{\partial N\over \partial x^\mu(\sigma)}
2471: {\partial M \over \partial \pi_\mu(\sigma)}\right),
2472: \labeq{pb}
2473: \ee
2474: where $\left(\partial x^\mu(\sigma') / \partial x^\nu(\sigma)\right)
2475: = \left(\partial \pi_\nu (\sigma') / \partial \pi_\mu (\sigma)\right)
2476: = \delta^\mu_\nu \delta^p(\sigma-\sigma')$, with other partial
2477: derivatives being zero.
2478:
2479: One way to calculate the Poisson brackets between
2480: a set of constraints $C$ and $C_i$,
2481: is to start from a putative Hamiltonian
2482: \be
2483: H=\int d^p \sigma ({A\over 2} C+ A^iC_i),
2484: \labeq{hame}
2485: \ee
2486: where $A$ and $A^i$ are arbitrary functions of $\sigma$, and
2487: then compute
2488: Hamilton's equations for the $\tau$ derivatives of
2489: $x^\mu$ and $\pi_\mu$. We then
2490: use these to
2491: determine the corresponding $\tau$ derivatives of
2492: $C$ and $C_i$. Setting these equal to
2493: $\{C,H\}$ and
2494: $\{C_i,H\}$ with $H$ given by (\ref{eq:hame}),
2495: we are able to infer the Poisson brackets between the constraints.
2496: For the Hamiltonian (\ref{eq:hame}) with $C$ and
2497: $C_i$ given in (\ref{eq:secondcc}), Hamilton's equations read
2498: \ba
2499: \dot{x}^\mu &=& A g^{\mu \nu} \pi_\nu + A^i x^\mu_{\,,i}\cr
2500: \dot{\pi}_\mu &=& (A^i \pi_\mu )_{,i} - {1\over 2} A g^{\lambda \nu}_{\quad,\mu}
2501: \pi_\lambda
2502: \pi_\nu + \mu_p^2 (A \overline{\gamma}\, \overline{\gamma}
2503: ^{ij} x^\nu_{,j})_{,i} g_{\mu \nu}
2504: + \mu_p^2 A g_{\mu \nu} \Gamma^{\nu}_{\lambda \epsilon}
2505: \overline{\gamma}\, \overline{\gamma}^{ij} x^\lambda_{,i}x^\epsilon_{,j},
2506: \labeq{hameqsa}
2507: \ea
2508: where dots denote $\tau$ derivatives and
2509: $\overline{\gamma}_{ij} ={x}^\mu_{,i} x^\nu_{,j} g_{\mu \nu}$ is the
2510: induced spatial metric on the brane and $ \overline{\gamma}$ its
2511: determinant. We have made use of the formula
2512: $d \overline{\gamma}= \overline{\gamma} \,\overline{\gamma}^{ij}
2513: d\overline{\gamma}_{ij}$.
2514:
2515: Using (\ref{eq:hameqsa}), it is a matter of
2516: straightforward algebra to compute $\dot{C}$ and
2517: $\dot{C_i}$ and hence infer all of
2518: the Poisson brackets. We find
2519: \ba
2520: \{C(\sigma),C(\sigma')\}&=&
2521: \left[ (8 \mu_p^2 \overline{\gamma}\, \overline{\gamma}^{ij}
2522: C_j)(\sigma){\partial
2523: \over \partial \sigma^i} + 4 \mu_p^2 (
2524: \overline{\gamma}\, \overline{\gamma}^{ij} C_j)_{,i}(\sigma) \right]\delta^p
2525: (\sigma-\sigma') \cr
2526: \{C(\sigma),C_i(\sigma')\}&=&
2527: \left[ 2 C(\sigma) {\partial
2528: \over \partial \sigma^i} +C_{,i}(\sigma)\right]
2529: \delta^p (\sigma-\sigma') \cr
2530: \{C_i(\sigma),C_j(\sigma')\}&=&
2531: \left[
2532: C_i(\sigma) {\partial
2533: \over \partial \sigma^j}
2534: +C_j(\sigma) {\partial
2535: \over \partial \sigma^i}
2536: +{\partial C_{i}
2537: \over \partial \sigma^j}(\sigma)\right]
2538: \delta^p (\sigma-\sigma').
2539: \labeq{results}
2540: \ea
2541: The right hand side consists of linear combinations
2542: of the constraints and, hence, it vanishes weakly.
2543: We conclude that the constraint algebra closes and,
2544: hence that the constraints are first class. Notice that
2545: the case of strings, $p=1$, is specially simple since
2546: $\overline{\gamma}\, \overline{\gamma}^{11}= 1$ and the
2547: Poisson bracket algebra is linear, with field-independent
2548: structure constants.
2549:
2550: The calculation also provides
2551: a consistency check on our Hamiltonian (\ref{eq:totalh1}),
2552: which is precisely of the form (\ref{eq:hame}), since it implies
2553: the constraints are preserved under
2554: time evolution in $\tau$.
2555:
2556: \section*{Appendix 3: Equivalence of gauge-fixed Hamiltonian and Lagrangian
2557: equations}
2558:
2559: In this Appendix we establish that
2560: the Lagrangian
2561: equations following from the gauge-fixed action (\ref{eq:newact})
2562: for winding $M2$-branes
2563: are equivalent to the Lagrangian equations for
2564: a string in the time-dependent background
2565: $g_{\mu \nu} = |t| \eta_{\mu \nu}$, in a certain string
2566: worldsheet coordinate system. This is in accord
2567: with our general arguments.
2568:
2569: The equations of motion
2570: following from the gauge-fixed action (\ref{eq:newact})
2571: are:
2572: \ba
2573: \ddot{\vec{x}} &=& t^2 \vec{x}\,''+ 2 t t' \vec{x}\,'\cr
2574: \ddot{t} &=& t \vec{x}\,'\,^2+ t {t'}^2 + t''t^2.
2575: \labeq{new}
2576: \ea
2577: and the constraints take the form
2578: \be
2579: \dot{t} t'=\dot{\vec{x}}\cdot\vec{x}\,'; \qquad \dot{t}^2= \dot{\vec{x}}\,^2
2580: + t^2 (\vec{x}'\,^2 -{t'}^2).
2581: \labeq{constrs}
2582: \ee
2583:
2584:
2585: We want to compare these equations with the Lagrangian
2586: equations of motion following
2587: from the Polyakov action (\ref{eq:pact}), with $p=1$. These are
2588: \ba
2589: &&\partial_\tau((-\gamma)^{1\over 2}\gamma^{\tau \tau} \partial_\tau x^\mu)
2590: +\partial_\sigma((-\gamma)^{1\over 2} \gamma^{\sigma \sigma}
2591: \partial_\sigma x^\mu)\cr
2592: &&+(-\gamma)^{1\over 2}
2593: \Gamma_{\nu \lambda}^\mu (\gamma^{\tau \tau}\partial_\tau x^\nu\partial_\tau
2594: x^\lambda+\gamma^{\sigma \sigma}
2595: \partial_\sigma x^\nu
2596: \partial_\sigma x^\lambda)=0,
2597: \labeq{strbg}
2598: \ea
2599: where $\Gamma_{\nu \lambda}^\mu$ is the Christoffel symbol for
2600: the background metric.
2601:
2602: We also have the
2603: constraints that the worldsheet metric $\gamma_{\alpha \beta}$
2604: is conformal to the induced metric on the string.
2605: We have the freedom to choose worldsheet coordinates
2606: on the string, but since the equations are conformally
2607: invariant, only the conformal class matters.
2608: The choice $\gamma_{\alpha \beta}
2609: = \Omega^2 {\rm diag}(-t^2, 1)$ is found
2610: to yield the two constraints (\ref{eq:constrs}).
2611:
2612: For our background, $g_{\mu \nu} = |t| \eta_{\mu \nu}$,
2613: we have nonzero Christoffel symbols
2614: $\Gamma^{0}_{00} = 1/(2t)$, $\Gamma^{i}_{j0} =
2615: \Gamma^{i}_{0j}= \delta_{ij} /(2t)$,
2616: $\Gamma^{0}_{ij} = \delta_{ij}/(2t)$, where $i$ runs over
2617: the background spatial indices $1$ to $d-1$.
2618: The string
2619: equations of motion (\ref{eq:strbg}) are then found
2620: to be equivalent to (\ref{eq:new}), for
2621: all nonzero $t$.
2622:
2623: From the string point of view, this choice of gauge
2624: would seem arbitrary, and indeed it would appear
2625: to be degenerate at $t=0$. Yet, as we have seen,
2626: this gauge choice is just $A=1$ and $A^i=0$, which is
2627: entirely natural from the canonical point of view.
2628: It has the desirable property that
2629: the equations of motion and the constraints are
2630: regular at $t=0$, and from the general properties of
2631: the canonical formalism we are guaranteed the
2632: existence of an infinite class of coordinate
2633: systems, related by nonsingular
2634: coordinate transformations,
2635: in which the equations of motion will remain regular.
2636:
2637:
2638: \section*{Appendix 4: Ordering ambiguities and their resolution
2639: for relativistic
2640: particles}
2641:
2642: In the main text we have discussed the canonical Hamiltonian
2643: treatment of relativistic particles and $p$-branes. When
2644: one comes to quantize these theories in a general
2645: background, certain ordering ambiguities appear
2646: which must be resolved. Here we provide a brief overview,
2647: following the
2648: more comprehensive discussion in Ref. \cite{dewitt}.
2649:
2650: The field equation for a relativistic particle is simply the
2651: expression of the quantum Hamiltonian constraint $H= 0$, in
2652: a coordinate space representation. The first
2653: task is to determine the representation of the momentum operator
2654: $p_\mu$ in this representation, and then that of the Hamiltonian
2655: operator $H$. As we shall now discuss, this requires
2656: knowledge of
2657: the metric on the space of coordinates. We shall only
2658: deal with the point particle case.
2659:
2660: The classical Hamiltonian constraint
2661: for a massive particle
2662: in a background metric $g_{\mu \nu}$ reads
2663: \be
2664: g^{\mu \nu} p_\mu p_\nu + m^2 \approx 0.
2665: \labeq{classpart}
2666: \ee
2667: First we attempt to determine the coordinate space
2668: representation of $p_\mu$, consistent
2669: with the quantum bracket:
2670: \be
2671: \left[ x^\mu, p_\nu\right] =i \hbar \delta_\nu^\mu.
2672: \labeq{braq}
2673: \ee
2674: One choice is $p_\mu = -i\hbar \partial_\mu$
2675: but this is not unique:
2676: the representation $p_\mu = -i\hbar(\partial_\mu +f_\mu)$,
2677: with $f_\mu$ any function of the coordinates $x^\mu$ and $\tau$,
2678: is equally good as far as (\ref{eq:braq}) is concerned.
2679:
2680: We now show how $f_\mu$ may be determined from
2681: the additional requirement that $p_\mu$ be hermitian {\it i.e.,}
2682: that the momentum be real.
2683: In the coordinate space representation, this requirement reads
2684: \be
2685: \langle \chi| p_\mu |\phi\rangle
2686: \equiv \int d^d x (-g(x))^{1\over 2} \left( \chi^* p_\mu \phi\right)
2687: = \langle \phi| p_\mu |\chi \rangle^* \
2688: \equiv \int d^d x (-g(x))^{1\over 2} \left( \phi^* p_\mu \chi\right)^*
2689: \labeq{ip}
2690: \ee
2691: where the integration runs over the space of coordinates
2692: and $g_{\mu \nu}$ is the metric on that space.
2693: It is straightforward to check that
2694: the naive operator $-i\hbar \partial_\mu$ is in fact {\it not}
2695: hermitian for general $g_{\mu \nu}$, but that
2696: \be
2697: p_\mu = -i\hbar \left (\partial_\mu + {1\over 4} \left(\partial_\mu \ln (-
2698: g)\right)\right) =
2699: -i\hbar (-g)^{-{1\over 4}} \partial_\mu (-g)^{1\over 4}
2700: \labeq{moment}
2701: \ee
2702: is. This discussion uniquely determines the real part of $f_\mu$:
2703: an imaginary part may be absorbed in
2704: an unobservable phase of the wavefunction\cite{dewitt}.
2705:
2706: Similarly, when we consider the Hamiltonian constraint (\ref{eq:classpart}),
2707: the questions arise of where to place the $g^{\mu \nu}$ relative
2708: to the $p_\mu$'s, and whether to include
2709: any factors of the metric determinant $g$. The resolution
2710: is familiar: if we write the Hamiltonian in covariant
2711: derivatives on the space of coordinates, it will be hermitian
2712: since we can integrate by parts ignoring the $\sqrt{-g}$
2713: factor in the measure. This suggests setting the first term in
2714: (\ref{eq:classpart}) equal to the scalar Laplacian:
2715: \be
2716: g^{\mu \nu} p_\mu p_\nu \rightarrow -\hbar^2 (-g)^{-{1\over 2}}\partial_\mu (-
2717: g)^{1\over 2} g^{\mu \nu}
2718: \partial_\mu = (-g)^{-{1\over 4}} p_\mu
2719: (-g)^{1\over 2} g^{\mu \nu} p_\nu (-g)^{-{1\over 4}}.
2720: \labeq{orderh}
2721: \ee
2722: It is straightforward to check that this is the only
2723: choice of ordering and powers of $(-g)$ which is hermitian and
2724: has the correct classical limit.
2725: Nevertheless, this
2726: ordering is not immediately apparent!
2727: More generally,
2728: one can also include
2729: terms involving commutators of $p_\mu$ which
2730: are zero in the classical limit, but which
2731: produce the Ricci scalar $R$ in the quantum
2732: Hamiltonian\cite{dewitt}. In the space-time we consider
2733: $R$ is zero. Hence, such
2734: terms do not arise.
2735:
2736:
2737:
2738: \begin{thebibliography}{9999}
2739: \bibitem{kosst}
2740: J.~Khoury, B.A.~Ovrut, N.~Seiberg, P.J.~Steinhardt
2741: and N.~Turok, Phys. Rev. {\bf D65}, 086007 (2002).
2742: \bibitem{steif}
2743: G.T.~Horowitz and A.R.~Steif, Phys. Rev. {\bf D42} (1990) 1950; Phys. Lett. {\bf
2744: B258} (1991) 91.
2745: \bibitem{tolley1} A.J.~Tolley and N.~Turok, Phys. Rev. {\bf D66} (2002) 106005.
2746:
2747: \bibitem{ekperts}
2748: J.~Khoury, B.~A. Ovrut, P.~J. Steinhardt, and N.~Turok,
2749: Phys. Rev. {\bf D66}, 046005 (2002).
2750: \bibitem{greene}
2751: P.~S.~Aspinwall, B.~R.~Greene and D.~R.~Morrison,
2752: hep-th/9309097,
2753: Nucl.\ Phys.\ B {\bf 416}, 414 (1994); E.~Witten, hep-th/9301042,
2754: Nucl.\ Phys.\ B {\bf 403}, 159 (1993).
2755: \bibitem{kost}
2756: J.~Khoury, B.A.~Ovrut, P.J.~Steinhardt, and N.~Turok,
2757: Phys. Rev. {\bf D64}, 123522 (2001).
2758: \bibitem{STu}
2759: P.J.~Steinhardt and N.~Turok, Science {\bf 296}, 1436 (2002); Phys.
2760: Rev. {\bf D65}, 126003 (2002).
2761: \bibitem{review}
2762: P.J.~Steinhardt and N.~Turok, astro-ph/0404480.
2763: \bibitem{chaos}
2764: J.K.~Erickson, D.H.~Wesley, P.J.~Steinhardt, and N.~Turok,
2765: Phys. Rev. {\bf D69}, 063514 (2004).
2766: \bibitem{hw}
2767: P.~Ho\v rava and E.~Witten,
2768: Nucl. Phys. {\bf B460} (1996) 506; {\bf B475} (1996) 94.
2769: \bibitem{ovrut} A.~Lukas, B.A.~Ovrut and D.~Waldram,
2770: Nucl. Phys. {\bf B495} (1997) 365.
2771: \bibitem{tolleyperts}
2772: A.J. Tolley, N. Turok and P.~J. Steinhardt,
2773: Phys. Rev. {\bf D69}, 106005 (2004).
2774: \bibitem{seibergetal} H.~Liu, G.~Moore and N.~Seiberg, JHEP {\bf 0206} (2002) 045;
2775: JHEP {\bf 0210} (2002) 031; O.~Aharony, M.~Fabinger, G.~Horowitz and E.~
2776: Silverstein,
2777: hep-th/0204158; M.~Fabinger and J.~McGreevy, hep-th/0206196;
2778: M.~Fabinger and S.~Hellerman, hep-th/0212223; G.T.~Horowitz and J.~Polchinski,
2779: Phys. Rev. {\bf D66} (2002) 103512; L.~Cornalba and M.S.~Costa, Phys. Rev. {\bf
2780: D66} (2002) 066001;
2781: L.~Cornalba, M.S.~Costa and C.~Kounnas, Nucl. Phys. {\bf B637} (2002) 378;
2782: L.~Cornalba and M.S.~Costa, hep-th/0302137.
2783: \bibitem{ccosta}
2784: L.~Cornalba and M.S.~Costa, hep-th/0310099.
2785: \bibitem{dewitt} B.~S. DeWitt, Rev. Mod. Phys. {\bf 29}, 377 (1957);
2786: in {\it Relativity, Groups and Topology},
2787: Les Houches lectures (1963), Gordon and Breach, New York, 1964.
2788: \bibitem{bachas} C.~Bachas, Phys. Lett. {\bf B374} 37 (1996), hep-th/9511043.
2789: \bibitem{pioline}
2790: M.~Berkooz and B.~Pioline, JCAP 0311 (2003) 007, hep-th/0307280; M.~Berkooz,
2791: B.~Pioline and M.~Rozali,
2792: hep-th/0405126.
2793: \bibitem{misgive} N.~Turok, in
2794: {\it The Future of Theoretical Physics and Cosmology : Celebrating Stephen
2795: Hawking's 60th Birthday}, eds. G.W.~Gibbons, E.P.S.~Shellard and S.J.~Rankin,
2796: Cambridge University Press, 2003.
2797: \bibitem{dirac} P.A.M. Dirac, {\it Lectures on Quantum Mechanics}, Belfer
2798: Graduate School of Science Monographs Series, New York, 1964.
2799: \bibitem{AS}
2800: M. Abramowitz and I.A. Stegun, {\it Handbook of Mathematical
2801: Functions}, Dover, New York, 1970, page 686.
2802: \bibitem{birrell} See e.g. N.~D. Birrell and P.~C.~W. Davies, {Quantum Fields
2803: in Curved Space-time}, CUP, 1982.
2804: \bibitem{witten} E.~Witten, Nucl. Phys. {\bf B471} (1996) 135;
2805: Nucl. Phys. {\bf B443} (1995) 85.
2806: \bibitem{devega} H.J. de Vega,
2807: I. Giannakis, A.
2808: Nicolaidis, Mod. Phys. Lett. {\bf A10} 2479 (1995); I. Antoniadis and
2809: G. Savvidy, hep-th/0402077 and references therein.
2810: \bibitem{nicolaidis} A. Nicolaidis, J.E. Paschalis,
2811: P.I. Porfyriadis,
2812: Phys. Rev. {\bf D58} 047901 (1998);
2813: hep-th/9702185
2814: \bibitem{maeda}
2815: G.W. Gibbons and K. Maeda, Nucl. Phys. {\bf B298} 741 (1988).
2816: \bibitem{duff}
2817: M.J. Duff, TASI Lectures, hep-th/9912164.
2818: \bibitem{heading}
2819: J. Heading, {\it Phase Integral Methods}, Methuen Physical Monographs,
2820: London, 1962, page 85.
2821: \bibitem{martinec}
2822: A.~E. Lawrence and E. Martinec, Class. Quant. Grav. {\bf 13} (1996)
2823: 63; hep-th/9509149.
2824: \bibitem{gubser} S.~S. Gubser, hep-th/0305099.
2825: \bibitem{coleman}
2826: K. Lee, Phys. Rev. Lett., {\bf 61} (1988) 263; S. Coleman and
2827: K. Lee, Nuc. Phys. {\bf B329} (1990) 387.
2828: \bibitem{bubbles}
2829: S. Coleman, Phys. Rev. {\bf D15}, 2929 (1977); C. G. Callan Jr.
2830: and S. Coleman, Phys. Rev. {\bf D16}, 1762 (1977).
2831: \bibitem{capovilla} R. Capovilla, J. Guven and E. Rojas, hep-th/0404178;
2832: Nucl. Phys. Proc. Suppl. {\bf 88} 337 (2000).
2833: \bibitem{townsend} J.A.~Azcarraga, J.M.~Izquierdo and P.K.~Townsend,
2834: Phys. Lett. {\bf B267} 366 (1991); Phys. Rev. {\bf D45} 3321 (1992).
2835: \bibitem{gutowski} J.~Gutowski, G.~Papadopoulos and P.K.~Townsend,
2836: Phys.Rev. {\bf D60} 106006 (1999), hep-th/9905156.
2837: \end{thebibliography}
2838: \end{document}
2839:
2840:
2841:
2842:
2843:
2844:
2845: