1: \documentclass[12pt,letterpaper]{JHEP3}
2: \usepackage{cite}
3: \usepackage{epsfig}
4: %\usepackage{amsfonts}
5: %\usepackage{showkeys}
6:
7:
8: \title{Precision cosmology and the landscape}
9:
10: \author{%
11: Raphael Bousso\\
12: Center for Theoretical Physics, Department of Physics\\
13: University of California, Berkeley, CA 94720-7300, U.S.A.\\
14: {\em and}\\
15: Lawrence Berkeley National Laboratory,
16: Berkeley, CA 94720-8162, U.S.A.}
17:
18: \abstract{%
19: %
20: After reviewing the cosmological constant problem---why is $\Lambda$
21: not huge?---I outline the two basic approaches that had emerged by
22: the late 1980s, and note that each made a clear prediction.
23: Precision cosmological experiments now indicate that the
24: cosmological constant is nonzero. This result strongly favors the
25: environmental approach, in which vacuum energy can vary discretely
26: among widely separated regions in the universe. The need to explain
27: this variation from first principles constitutes an observational
28: constraint on fundamental theory. I review arguments that string
29: theory satisfies this constraint, as it contains a dense discretuum
30: of metastable vacua. The enormous landscape of vacua calls for
31: novel, statistical methods of deriving predictions, and it prompts
32: us to reexamine our description of spacetime on the largest scales.
33: I discuss the effects of cosmological dynamics, and I speculate that
34: weighting vacua by their entropy production may allow for prior-free
35: predictions that do not resort to explicitly anthropic arguments.
36: %
37: }
38:
39: \preprint{\hepth{0610211}}
40:
41: \begin{document}
42:
43: \section{Introduction}
44:
45: The quest for quantum gravity is driven by a desire for consistency
46: and unity of physical law. Quantum mechanics and the general theory
47: of relativity are hard to fit under one roof. String theory succeeds
48: at this task, exhibiting a level of mathematical rigor and richness of
49: structure that has yet to be matched by other approaches.
50:
51: Unfortunately, the subject has been lacking guidance from experiment.
52: Particle accelerators, in particular, are unlikely to probe effects of
53: quantum gravity directly. The energies that can be attained are many
54: orders of magnitude too low. This problem has nothing to do with
55: string theory. It arises the minute we turn our attention to quantum
56: gravity, because gravity is extremely weak in scattering experiments.
57:
58: On large scales, however, gravity rules. The expansion of the
59: universe dominates over all other dynamics at distances above 100 Mpc.
60: Similarly, once matter condenses enough to form a black hole, no known
61: force can prevent its total collapse into a singularity. In the early
62: universe, moreover, quantum effects can be important. Perhaps, then,
63: string theory should be looking towards cosmology for guidance.
64:
65: In fact, recent years have seen a remarkable transformation. String
66: theory has become driven, to a significant extent, by the results of
67: precision experiments in cosmology. The discovery of dark
68: energy~\cite{Per98,Rie98} suggests that vacuum energy is an
69: environmental variable. String theory naturally provides for
70: variability of the cosmological constant, with a fine enough spacing
71: to accommodate the observed value~\cite{BP}. In this sense, recent
72: cosmological observations constitute observational evidence for the
73: theory. Moreover, they have focussed attention on the large number of
74: metastable vacua---the string
75: landscape~\cite{BP,KKLT,Sus03}---believed to be responsible for this
76: variability.\footnote{For a detailed review and extensive references,
77: see Ref.~\cite{DouKac06}. For a less technical discussion of the
78: issues covered in the present article, see Ref.~\cite{BouPol04}.}
79:
80: Before presenting conclusions from precision experiments, I will argue
81: in Sec.~\ref{sec-why} that much can be learned from far more primitive
82: observations of the cosmos. A simple question---why is the universe
83: so large?---translates into a number of major challenges to
84: theoretical cosmology. One, the flatness problem, motivated the
85: theory of inflation, which went on to explain the origin of structure
86: in the universe, making a number of specific predictions. Another,
87: the cosmological constant problem, is especially closely related to
88: fundamental theory: Why is the energy of empty space more than 120
89: orders of magnitude smaller than predicted by quantum field theory?
90:
91: Most early discussions of the cosmological constant problem tended to
92: embrace one of two distinct approaches. Either the cosmological
93: constant has to be zero due to some unknown symmetry; or it is an
94: environmental variable that can vary over distances large compared to
95: the visible universe, and observers can only live in regions where it
96: is anomalously small. Though neither of these approaches had been
97: developed into concrete models, each made a signature prediction: that
98: the cosmological constant is zero, or that it is small but nonzero.
99:
100: The refined experiments of the last ten years have amassed additional
101: evidence for inflation, and they have managed to discriminate clearly
102: between the two approaches to the cosmological constant problem. The
103: discovery of nonzero dark energy, in particular, is precisely what the
104: second, environmental approach predicted, and it all but rules out the
105: first approach. These results, summarized in Sec.~\ref{sec-pc}, are
106: the empirical foundation of the landscape of string theory.
107:
108: In Sec.~\ref{sec-bp}, I will describe a concrete model that realizes
109: the second approach in string theory. The topological complexity of
110: compact extra dimensions leads to an exponentially large potential
111: landscape. Its metastable vacua form a dense ``discretuum'' of values
112: of the cosmological constant. Every vacuum will be realized in
113: separate regions, each bigger than the visible universe, but structure
114: (and thus observers) form only in those regions where the cosmological
115: constant is sufficiently small.
116:
117: In Sec.~\ref{sec-predict}, I will discuss some of the novel challenges
118: posed by the string landscape. The greatest challenge, perhaps, is to
119: develop methods for making predictions in a theory with $10^{500}$
120: metastable vacua. In fact, this difficulty is sometimes presented as
121: insurmountable, but I will argue that it just comes down to a lot of
122: hard work. In particular, I will argue that the correct statistical
123: treatment of vacua necessitates a departure from the traditional,
124: global description of spacetime. I will further propose a statistical
125: weighting of vacua based on entropy production, which performs well in
126: comparison with far more specific anthropic conditions. A general
127: weighting of this type may pave the way for a calculation of the size
128: of the universe from first principles.
129:
130:
131:
132: \section{Why is the universe large?}
133: \label{sec-why}
134:
135: In cosmology, the most naive questions can be the most profound. A
136: famous example is Olbers' paradox: Why is the sky dark at night? In
137: this spirit, let us ask why the universe is large. To quantify
138: ``large'', recall that only a single length scale can be constructed
139: from the known constants of nature: the Planck length
140: \begin{equation}
141: l_{\rm P} = \sqrt{\frac{G\hbar}{c^3}}
142: \approx 1.616 \times 10^{-33} {\rm cm}~.
143: \end{equation}
144: Here $G$ denotes Newton's constant and $c$ is the speed of
145: light.\footnote{In the remainder I will work mostly in Planck
146: units. For example, $t_{\rm P}=l_{\rm P} /c \approx .539 \times
147: 10^{-43} {\rm s}$ and $M_{\rm P} = 2.177\times 10^{-5} {\rm g}$.}
148:
149: The actual size of the universe is larger than this fundamental length
150: by a factor
151: \begin{equation}
152: H^{-1} = .8 \times 10^{61}~.
153: \label{eq-h}
154: \end{equation}
155: Here $H\approx 70$ km/s/Mpc is the Hubble scale, and $H^{-1}$ is the
156: Hubble length. Of course, this refers to the size of the universe as
157: we see it today, and thus is only a lower bound on the length scales
158: that may characterize the universe as a whole.\footnote{As I shall
159: discuss below, there is evidence that the universe is exponentially
160: larger than the visible universe, but that we will never see a
161: region larger than $.98\times 10^{61}$, no matter how long we
162: wait.}
163:
164: The dynamical behavior of a system usually reflects the scales of the
165: input parameters, and other scales constructed from them by
166: dimensional analysis. For example, the ground state of a harmonic
167: oscillator of mass $m$ and frequency $\omega$ has a position
168: uncertainty of order $(m\omega)^{-1/2}$. The parameters entering
169: cosmology are $G$, $\hbar$, and $c$, so $l_{\rm P}$ is the natural
170: length scale obtained by dimensional analysis. Thus, Eq.~(\ref{eq-h})
171: represents an enormous hierarchy of scales. Where does this large
172: number come from?
173:
174: At the very end of this article, I will speculate about the origin of
175: the number $10^{61}$. For now, let us simply consider the qualitative
176: fact that the universe is large compared to the Planck scale---a fact
177: that is plain to the naked eye, no precision experiments required. We
178: will see that some of the most famous problems in theoretical
179: cosmology are tied to this basic observation: the flatness problem,
180: and the cosmological constant problem.
181:
182: \subsection{The flatness problem and inflation}
183: \label{sec-inflation}
184:
185: We live in a universe that is spatially isotropic and homogeneous on
186: sufficiently large scales. The spatial curvature is constant, and it
187: is remarkably small. By the Einstein equation, this can be related to
188: the statement that the average density $\rho$ is not far from the
189: critical density,
190: \begin{equation}
191: \Omega\equiv \frac{\rho}{ \rho_{\rm c}} \sim O(1)~,
192: \end{equation}
193: where
194: \begin{equation}
195: \rho_{\rm c}= \frac{3H^2}{8\pi} ~.
196: \end{equation}
197:
198: This is surprising because it means that the early universe was flat
199: to fantastic accuracy. Through much of the history of the universe,
200: $\Omega$ has been pushed away from $1$. Einstein's equation implies
201: that
202: \begin{equation}
203: |\Omega-1| = (\dot{a})^{-2}~,
204: \label{eq-omo}
205: \end{equation}
206: where $\dot a$ is the time derivative of the scale factor of the
207: universe. The early universe was dominated by radiation for some
208: 70,000 years, and $a$ was proportional to $t^{1/2}$. Afterwards it
209: was dominated by matter for several billion years, with $a\propto
210: t^{2/3}$. Curvature would have become dominant ($\Omega\neq O(1)$)
211: over this time unless
212: \begin{equation}
213: |\Omega-1| \lesssim 10^{-59}
214: \end{equation}
215: when the universe began. This is the flatness problem.
216:
217: The flatness problem is closely related to our original question:
218: without flatness, the universe could would not have become large.
219: Suppose, for example, that the early universe had been tuned to
220: flatness less precisely, say $\Omega=1+10^{-20}$. This would have
221: been a closed universe, which would have expanded to a maximum radius
222: $10^{20}$ and recollapsed in a big crunch, all within about a time
223: $10^{20}$. In other words, this universe would have grown no larger
224: than a proton and lived for less than $10^{-23}$ seconds.
225:
226: If the universe had started out slightly underdense (say,
227: $\Omega=1-10^{-20}$), it would have developed a noticably ``open'',
228: i.e., hyperbolic spatial geometry after $10^{-23}$ seconds, when the
229: largest structures were the size of a proton. After this time,
230: density perturbations would no longer grow and structure formation
231: would cease. The largest coherent structures, each the size of a
232: proton, would freely stream apart. There would be no objects
233: comparable to the size of a planet, let alone galaxies. In this
234: sense, the universe would be small.
235:
236: A solution to the flatness problem appeared in the early 1980's:
237: inflation. (It simultaneously addressed a number of other major
238: conundra, such as the horizon problem.) For a detailed treatment,
239: see, e.g., Refs.~\cite{LidLyt93,LidLyt00}.
240:
241: The idea is to use Eq.~(\ref{eq-omo}) to our advantage: if $\dot a$
242: increases with time, then $\Omega$ is driven to $1$. This can be
243: accomplished by positing that the very early universe was dominated by
244: the vacuum energy of a scalar field before yielding to the standard
245: radiation era. The scale factor grows almost exponentially with time,
246: and $\Omega$ quickly approaches $1$ with exponential accuracy:
247: \begin{equation}
248: |\Omega-1| \approx e^{-2N}~,
249: \end{equation}
250: Here $N$ is the number of $e$-foldings, i.e., $e^N$ is the ratio
251: between the scale factor before and after inflation.
252: %
253: \EPSFIGURE{fig-visible.eps,width=8cm}{\label{fig-visible} If the early
254: universe underwent inflation, the universe may well be exponentially
255: larger than the visible portion (the interior of our past
256: lightcone).}
257: %
258:
259: Depending on the energy scale at which inflation occurred, perhaps 60
260: $e$-foldings suffice to guarantee that $.1\leq\Omega\leq 2$ today.
261: But it is easy to write down inflationary models with thousands or
262: millions of $e$-foldings. In such models, the universe would be spatially
263: flat not only on the present horizon scale, but on exponentially
264: larger scales, which will become visible only after an exponentially
265: longer time than the 13 billion years that have elapsed since the end
266: of inflation.
267:
268: The true abundance of such models in the potential landscape we get
269: from fundamental theory (Sec.~\ref{sec-predict}) is not yet known.
270: But apparently it is not exceedingly hard to get 60 $e$-foldings, or
271: else we would have seen curvature long ago. This suggests that models
272: with more $e$-foldings are not very rare. It would seem to require
273: some tuning for inflation to have lasted just long enough for the
274: first observable deviations from flatness to occur in the present era.
275: Thus, most inflationary theorists considered $\Omega=1$ to be a
276: prediction of inflation. By the same token, one would expect that the
277: universe is much larger than the visible universe, perhaps by as much
278: as $10^{100}$ or $10^{100000}$ (Fig.~\ref{fig-visible}).
279:
280:
281:
282: \subsection{The cosmological constant problem}
283: \label{sec-ccp}
284:
285: When Einstein wrote down the field equation for general relativity,
286: \begin{equation}
287: R_{\mu\nu} - \frac{1}{2} R g_{\mu\nu} + \Lambda g_{\mu\nu} =
288: 8\pi T_{\mu\nu}
289: \end{equation}
290: he had a choice: The cosmological constant $\Lambda$ was not fixed by
291: the structure of the theory. There was no formal reason to set it to
292: zero, and in fact, Einstein famously tuned it to yield a static
293: cosmological solution---his ``greatest blunder''.
294:
295: The universe has turned out not to be static, and $\Lambda$ was
296: henceforth assumed to vanish. This was never particularly satisfying
297: even from a classical perspective. The situation is not dissimilar to
298: a famous problem with Newtonian gravity---that there is no formal
299: necessity to equate the gravitational charge with inertial mass.
300:
301: In any case, the simple fact that the universe is large implies that
302: $|\Lambda|$ is small. I will show this first for the case of positive
303: $\Lambda$. Assume, for the sake of argument, that no matter is
304: present ($T_{\mu\nu}=0$). Then the only isotropic solution to
305: Einstein's equation is de~Sitter space, which exhibits a cosmological
306: horizon of radius
307: \begin{equation}
308: R_\Lambda= \sqrt{3/\Lambda}~.
309: \end{equation}
310: A cosmological horizon is the largest observable distance scale, and
311: the presence of matter will only decrease the horizon
312: radius~\cite{Bou00a}. We see scales that are large in Planck units,
313: so the cosmological constant must be small in these natural units.
314:
315: Negative $\Lambda$ causes the universe to recollapse independently of
316: spatial curvature, on a timescale of order $\Lambda^{-1/2}$. The
317: obvious fact that the universe is old compared to the Planck time then
318: implies that $|\Lambda|$ is small.
319:
320: These qualitative conclusions do not require any careful measurements.
321: Let us plug in some crude numbers that would have been available
322: already thirty years ago, such as the size of the horizon given in
323: Eq.~(\ref{eq-h}), or an age of the universe of order $10^{10}$
324: years. They imply that
325: \begin{equation}
326: |\Lambda|\lesssim 10^{-122}~.
327: \label{eq-small}
328: \end{equation}
329: Hence $\Lambda$ is very small indeed.
330:
331: This result makes it tempting to cast scruples aside and simply set
332: $\Lambda=0$. But from a modern perspective, to eliminate $\Lambda$ in
333: the classical Einstein equation is not only arbitrary, but futile.
334: $\Lambda$ returns through the back door, via quantum contributions to
335: the stress tensor, $\langle T_{\mu\nu}\rangle$. It is this effect
336: that makes the cosmological constant problem so notorious.\footnote{In
337: parts, our discussion will follow Refs.~\cite{Wei89,Car00}, where
338: more details and references can be found.}
339:
340: In quantum field theory, the vacuum is highly nontrivial. Every mode
341: of every field contributes a zero point energy to the energy density
342: of the vacuum (Fig.~\ref{fig-loop}a). The corresponding stress
343: tensor, by Lorentz invariance, must be proportional to the metric:
344: \begin{equation}
345: \langle T_{\mu\nu} \rangle = -\rho_\Lambda g_{\mu\nu} ~.
346: \label{eq-lrho}
347: \end{equation}
348: Though it appears on the right hand side of Einstein's equation,
349: vacuum energy has the form of a cosmological constant, with
350: $\Lambda=8\pi\rho_\Lambda$.\footnote{This is why the mystery of the
351: smallness of $\rho_\Lambda$ is usually referred to as the
352: cosmological constant problem. But it would be more appropriate to
353: call it the vacuum energy problem, since the quantum contributions
354: to the vacuum energy are what makes the problem especially hard.}
355: Its magnitude will depend on the cutoff.
356: %
357: \EPSFIGURE{fig-loop.eps,width=.8\textwidth}{\label{fig-loop} Some
358: contributions to vacuum energy. (a) Virtual particle-antiparticle
359: pairs (loops) gravitate. This is mandated by the equivalence
360: principle, and has been verified experimentally to a high degree of
361: accuracy~\cite{Pol06}. The vacuum of the standard model abounds with
362: such pairs and hence should gravitate enormously. (b) Symmetry
363: breaking in the early universe (e.g., the chiral and electroweak
364: symmetry) shifts the vacuum energy by amounts dozens of orders of
365: magnitude larger than the observed value.}
366: %
367:
368: For example, consider the electron, which is well understood at least
369: up to energies of order $M=$ 100 GeV. Dimensional analysis implies
370: that electron loops up to this cutoff contribute of order $(100$
371: GeV$)^4$ to the vacuum energy, or $10^{-68}$ in Planck units. Similar
372: contributions are expected from other fields. The real cutoff is
373: probably of order the supersymmetry breaking scale, giving at least a
374: TeV$^4\approx 10^{-64}$. It may be as high as the Planck scale, which
375: would yield $\Lambda$ of order unity. Thus, quantum field theory
376: predicts $\Lambda$ to be some 60 to 120 orders of magnitude larger
377: than the experimental bound, Eq.~(\ref{eq-small}).
378:
379: Additional contributions come from the potentials of scalar fields,
380: such as the potential giving rise to symmetry breaking in the
381: electroweak theory (Fig.~\ref{fig-loop}b). The vacuum energy of the
382: symmetric and the broken phase differ by approximately $(200$
383: GeV$)^4$. Any other symmetry breaking mechanisms at higher or lower
384: energy (such as chiral symmetry breaking of QCD, $(300$ MeV$)^4$) will
385: also contribute.\footnote{Incidentally, this means that the vacuum
386: energy in the early universe was many orders of magnitude larger
387: than today. This follows from well-tested physics and has been
388: known for a long time, and it should have made us suspicious of the
389: idea that the vacuum energy somehow ``had'' to be exactly zero. If
390: it was ok to have lots of it a few billion years ago, what could be
391: fundamentally wrong with having some now? It also shows that any
392: mechanism that would set the vacuum energy to zero in the very early
393: universe cannot solve the cosmological constant problem, since
394: $|\Lambda|$ would become huge after symmetry breaking.}
395:
396: I have exhibited various unrelated contributions to the vacuum
397: energy. Each is dozens of orders of magnitude larger than the
398: empirical bound today, Eq.~(\ref{eq-small}). In particular, the
399: radiative correction terms from quantum fields are expected to be at
400: least of order $10^{-64}$. They can come with different signs, but it
401: would seem overwhelmingly unlikely for all of them to be carefully
402: arranged to cancel to such exquisite accuracy ($10^{-122}$) in the
403: present era.
404:
405: This is the cosmological constant problem: why is the vacuum energy
406: today so small? It represents an immense crisis in physics: a
407: discrepancy between theory and experiment, of 60 to 120 orders of
408: magnitude, in a quantity as basic as the weight of empty space.
409:
410:
411: \subsection{Strategies and predictions}
412:
413: Since the 1980s, various strategies for approaching the cosmological
414: constant problem have been suggested. They fall into two broad
415: classes, with each class facing chararacteristic challenges and making
416: a characteristic prediction. To give them a fair hearing, let us
417: assume the cosmological data available in the 1980s: the cosmological
418: constant is tightly bounded, but has not yet been measured directly.
419: It might vanish or it might not.
420:
421: \subsubsection{$\Lambda$ must vanish}
422: \label{sec-van}
423:
424: The first approach is to seek a universal symmetry principle that
425: requires that $\Lambda=0$ in our universe today. The problem, of
426: course, is that this challenge has yet to be met. (Supersymmetry
427: guarantees that radiative contributions to the cosmological constant
428: vanish, but in our universe supersymmetry is broken at a scale of at
429: least a TeV.) The challenge is not made easier by the fact that one
430: must allow for a large cosmological constant in the early universe,
431: when various symmetries were not yet broken.
432:
433: Assuming these challenges could be met, the first approach does make a
434: sharp prediction: $\Lambda=0$.
435:
436: \subsubsection{$\Lambda$ is variable}
437: \label{sec-var}
438:
439: The second strategy~\cite{Sak84,Ban85,Wei87} is to posit that the
440: universe is large---exponentially larger than the presently visible
441: portion---and that $\Lambda$ varies from place to place, though it can
442: be constant over very large distances. As I will explain below,
443: structure such as galaxies will only form in locations
444: where~\cite{Wei87}
445: \begin{equation}
446: -10^{-123}\lesssim\rho_\Lambda\lesssim 10^{-121}~.
447: \label{eq-weinberg}
448: \end{equation}
449: Since structure is presumably a prerequisite for the existence of
450: observers, we should then not be surprised to find ourselves in such a
451: region.
452:
453: Why is $\Lambda$ related to structure formation? To form galaxies and
454: clusters, the tiny density perturbations visible in the cosmic
455: microwave background radiation had to grow under their own gravity,
456: until they became non-linear and decoupled from the cosmological
457: expansion. This growth is logarithmic during radiation domination,
458: and linear in the scale factor during matter domination. Vacuum
459: energy does not get diluted so it inevitably comes to dominate the
460: energy density. As soon as this happens, perturbations cease to grow,
461: and the only structures that remain gravitationally bound are
462: overdense regions that have already gone nonlinear. This means that
463: there would be no structure in the universe if the cosmological
464: constant had been large enough to dominate the energy density before
465: the first galaxies formed~\cite{Wei87}. This leads to the upper bound
466: in Eq.~(\ref{eq-weinberg}). The lower bound comes about because the
467: universe would have recollapsed into a big crunch too rapidly if the
468: cosmological constant had been large and negative~\cite{BarTip}.
469:
470: The problem with the second strategy is twofold:
471: \begin{enumerate}
472: \item{It works only in a theory in which $\Lambda$ is a dynamical
473: variable whose possible values are sufficiently closely spaced
474: that Eq.~(\ref{eq-weinberg}) can be satisfied.}
475: \item{Assuming generic initial conditions, one would need to find a
476: mechanism by which at least one value of $\Lambda$ satisfying
477: Eq.~(\ref{eq-weinberg}) can be dynamically attained in a
478: sufficiently large region in the universe.}
479: \end{enumerate}
480: Supposing that these challenges can be met, one would expect our local
481: cosmological constant to be fairly typical among the possible values
482: of $\Lambda$ compatible with structure formation. In an evenly spaced
483: spectrum, most values of Lambda satisfying Eq.~(\ref{eq-weinberg})
484: will be of order $10^{-121}$; for example, only a very small fraction
485: will be of order $10^{-146}$.
486:
487: Thus, the ``variable $\Lambda$'' approach predicts~\cite{Wei87} that
488: the cosmological constant is not much smaller than required by
489: Eq.~(\ref{eq-weinberg}). This means that it will be large enough to
490: be detectable in the present era. In other words, the ``variable
491: $\Lambda$'' approach predicts that the vacuum energy should be nonzero
492: and comparable to the matter density today.
493:
494:
495: \section{Precision cosmology}
496: \label{sec-pc}
497:
498: Beginning with the measurement of anisotropies in the cosmic microwave
499: background in the 1990s~\cite{COBE}, experimental cosmology has
500: undergone a remarkable transformation. The subject has evolved from
501: order-of-magnitude estimates of a few cosmological parameters to
502: precise measurements of increasingly complex phenomena, leading to the
503: emergence of a ``standard model'' of cosmology. I will not attempt to
504: review these developments in any detail; see, e.g.,
505: Refs.~\cite{Teg03,CopSam06,Spe06,SelSlo06,Teg06}. Instead
506: I will summarize how several independent types of observations have
507: helped us evaluate the proposals discussed in the previous section.
508: This is shown schematically in Fig.~\ref{fig-precision}.
509: %
510: \EPSFIGURE{fig-precision.eps,width=.9\textwidth}{\label{fig-precision}
511: Recent cosmological precision data (light/red shading) strongly
512: support the idea that the cosmological constant is an environmental
513: variable that can scan densely spaced values. The thinner arrows
514: indicate that a result merely adds plausibility to another; the
515: thicker ones denote the most straightforward implication of a
516: result.}
517: %
518:
519: \subsection{Inflation looks good}
520: \label{sec-infgood}
521:
522: \begin{enumerate}
523:
524: \item{Measurements of fluctuations in the cosmic microwave background
525: radiation (CMB) strongly support inflation, in two ways:}
526:
527: \begin{itemize}
528: \item{The position of the peaks of the perturbation spectrum as a
529: function of angular scale imply that the universe is spatially flat
530: to excellent precision. Not only is $\Omega\sim O(1)$, but
531: $\Omega=1$, to accuracy of a few percent. This is the expected
532: result if inflation was the correct explanation of the flatness of
533: the universe.}
534: \item{The detailed spectrum of perturbations is nearly scale-invariant
535: and Gaussian. This is natural in inflationary models and rules
536: out many other possible seeds for structure formation, such as
537: topological defects.}
538: \end{itemize}
539:
540: \item{Measurements of the large scale structure (the distribution of
541: galaxies and galaxy clusters), by techniques such as weak lensing
542: and the Lyman alpha forest, are consistent with $\Omega=1$ and
543: have reduced the error bars on this result, supporting inflation.}
544:
545: \item{Supernova measurements have detected an extra contribution to
546: the total energy density in the universe, $\Omega_\Lambda=.7$.
547: Meanwhile, the observation of large scale structure has
548: corroborated the view that most pressureless matter is dark,
549: $\Omega_{\rm matter}=.3$. This implies independently of the
550: previous arguments that $\Omega=1$.}
551:
552: \end{enumerate}
553:
554: This evidence directly supports inflation. Thus, it indirectly lends
555: credence to the ``variable $\Lambda$'' approach the cosmological
556: constant problem (Sec.~\ref{sec-var}). That strategy requires that
557: the universe be much bigger than what we can presently see of it. As
558: I discussed at the end of Sec.~\ref{sec-inflation}, this type of
559: global picture is natural in inflationary theory.\footnote{In
560: Sec.~\ref{sec-predict} I will argue that one should not, in fact,
561: attempt to describe all of this global spacetime at once. Because
562: different regions are forever causally disconnected, they correspond
563: to different outcomes in a decoherent history.}
564:
565: \subsection{The cosmological constant is non-zero}
566: \label{sec-lnonzero}
567:
568: \begin{enumerate}
569:
570: \item{Supernova experiments show that the universe began accelerating
571: its expansion approximately seven billion years ago. This indicates
572: the presence of vacuum energy with $\rho_\Lambda=1.25\times
573: 10^{-123}$. Present data disfavor any time-dependence of this
574: component. Thus, the data strongly support the conclusion that the
575: cosmological constant is non-zero.}
576:
577: \item{The CMB and large scale structure measurements cited in
578: Sec.~\ref{sec-infgood} above reinforce this conclusion, since they
579: imply that $\Omega=1$. This value cannot be accounted for by dark
580: matter alone. It implies that at least $70\%$ of the universe
581: consists of energy that doesn't clump. The simplest such
582: component is a cosmological constant.}
583:
584: \item{Indirectly, this conclusion is also supported by the
585: measurements of the perturbation spectrum in the CMB cited above.
586: They favor inflationary models, and inflation generically predicts
587: $\Omega=1$.}
588:
589: \end{enumerate}
590:
591: In summary, there is now strong evidence that $\Lambda>0$. But a
592: non-zero value of $\Lambda$ in the observed range is precisely what
593: the ``variable $\Lambda$'' approach to the cosmological constant
594: problem (Sec.~\ref{sec-var}) predicted. This is rather fortunate,
595: since string theory naturally leads to a concrete implementation of
596: the ``variable $\Lambda$'' strategy, which I will discuss in
597: Sec.~\ref{sec-bp}.
598:
599: The data essentially rule out the ``$\Lambda$-must-vanish'' approach
600: (Sec.~\ref{sec-van}), since $\Lambda$ apparently does not vanish. But
601: one could argue that the approach has merely become less appealing,
602: requiring more epicycles to match observation. I will now try to
603: quantify this, before returning to the ``variable $\Lambda$''
604: strategy.
605:
606:
607: \subsection{The price of denial}
608: \label{sec-denial}
609:
610: The idea that $\Lambda$ is an environmental variable is a perfectly
611: logical possibility, but it does represent a retreat. An apt
612: analogy~\cite{Wei05} is Kepler's hope of explaining the relation
613: between planetary orbits from first principles. The hope was dashed
614: by Newton's theory of gravitation. Of course, that was no reason to
615: reject a theory of tremendous explanatory power. We simply came to
616: accept that the orbits are the results of historical accidents and
617: that there are many other solar systems in which different
618: possibilities are realized.
619:
620: But let us not be too hasty in abandoning the quest for a unique
621: prediction of today's value of $\Lambda$. Instead, let us ask what it
622: would take to maintain this type of approach in light of the discovery
623: of non-zero vacuum energy.
624:
625: We would need to assume that some symmetry or other effect makes
626: $\Lambda$ vanish, except for a correction of order $10^{-123}$. This
627: takes a miracle as the starting point: despite decades of work, no
628: mechanism has been found that requires $\Lambda=0$ without running
629: into conflict with known physics~\cite{Wei89,Pol06}. And supposing it
630: existed, how would any posited correction evade a mechanism so
631: powerful as to cancel out many enormous and disparate contributions to
632: vacuum energy (see Sec.~\ref{sec-ccp})? Finally, why does this
633: correction have just the right magnitude so as to be comparable to the
634: matter density at the present time?
635:
636: In short, the price of insisting on a unique prediction for the
637: cosmological constant is that the cosmological constant problem breaks
638: up into three problems, none of them solved:\footnote{In some
639: discussions, the cosmological constant problem is identified with
640: these three questions. But this implicitly assumes that $\Lambda$
641: is unique. Fundamentally, the cosmological constant problem is only
642: one question: why is the vacuum energy not huge? As I explained in
643: Sec.~\ref{sec-var}, the ``variable $\Lambda$'' approach predicts
644: that $\Lambda$ will be small, but large enough to be already
645: noticable in our era. Thus it avoids the first in our list of
646: questions; it answers the third before we have a chance to worry
647: about it; and the second question does not arise. Indeed, at
648: present it is senseless to ask why $\Lambda\neq 0$, since we know of
649: no reason why $\Lambda$ {\em should\/} vanish. That it is asked
650: anyway betrays only how deeply we had absorbed the prejudice that it
651: does.}
652: \begin{enumerate}
653: \item{What makes the cosmological constant vanish?}
654: \item{Why is the cosmological constant not exactly zero?}
655: \item{Why now?}
656: \end{enumerate}
657: The first of these three problems seems by far the hardest; in any
658: case, it has resisted several decades of attack. It is tempting to
659: assume it solved, and to speculate instead about the putative
660: correction that makes $\Lambda$ nonzero. But let us be mindful that
661: any results obtained in this manner will rest on wishful thinking.
662:
663: Among such approaches, dynamical scalar fields (``quintessence'') take
664: a prominent role, perhaps because they posit observable deviations
665: from the equation of state of a cosmological constant. I confess that
666: I find this development perplexing. Dynamical scalars do not match
667: the data better than a fixed cosmological constant, and they are
668: theoretically far more baroque.
669:
670: Scalar fields like to roll off to infinity rapidly, or quickly get
671: stuck in a local minimum. For a scalar to mimic vacuum energy and yet
672: exhibit nontrivial dynamics more than ten billion years after the big
673: bang, would require an extremely flat (but not exactly flat) potential
674: over an enormous range. This necessitates
675: tunings~\cite{KolLyt98,Car98,Wei00} that include, but go far beyond,
676: arbitrarily setting the present vacuum energy to a small value. Yet
677: further tuning~\cite{Car98} is needed to explain why the long-range
678: force associated with an almost massless scalar has not been
679: detected.\footnote{Some authors do confront these latter problems (see
680: Refs.~\cite{ChaHal04,Svr06} for recent examples). Aside from the
681: unsolved theoretical question of why $\Lambda$ should vanish at late
682: times, such models also receive increasing pressure from
683: observation, since dark energy does appear to be at least
684: approximately constant.}
685:
686: Thus, quintessence not only fails to address the very real question of
687: why $\Lambda$ is small, but, unprovoked by data, burdens us with the
688: challenge of explaining several additional very small numbers.
689:
690: Understandably, experimenters demand parametrizations of some spaces
691: of models that they can hope to constrain~\cite{Alb06}. But let us
692: not confuse models (which come cheaper the more complicated we make
693: them) with explanations. The danger is that we will forever abuse the
694: data to constrain ever more baroque models while overlooking the
695: simplest one~\cite{LidMuk06}.
696:
697: A cosmological constant is already favored by experiment, and it is
698: arguably the only model for which we have at least a tentative
699: fundamental explanation (Sec.~\ref{sec-bp}). If one finds this
700: explanation unattractive, it makes sense to seek a different origin of
701: the simplest model compatible with the data. What makes no sense is
702: to write down more complicated models than the data require, while
703: making no attempt to explain their origin in a credible fundamental
704: theory.\footnote{Similar remarks apply to the idea that gravity should
705: be modified to account for the apparent deviations from
706: $\Lambda=0$. This approach also makes sense only to the degree that
707: we have any reason to believe that $\Lambda$ should vanish at late
708: times, which we don't. In a modified gravity theory, the quantum
709: field theory contributions to the cosmological constant would be
710: just as large, unless one violates the equivalence principle, which
711: conflicts with other experiments~\cite{Pol06}.}
712:
713: I am not, of course, proposing that we stop looking experimentally for
714: any time dependence of dark energy. The evidence for a nonzero
715: cosmological constant is surely among the most profound insights ever
716: gained from experiment. This alone warrants every effort to confirm
717: and refine what we know about dark energy. Perhaps more surprises
718: await us, complicating the story further. Meanwhile, I feel that we
719: theorists would do well to solve the problems we actually have; those
720: are bad enough.
721:
722:
723: \section{The discretuum}
724: \label{sec-bp}
725:
726: I have argued that experiment favors the ``variable $\Lambda$''
727: approach to the cosmological constant problem. I have also
728: spelled out the main challenges to its implementation. In this
729: section, I will present evidence that these challenges are met by
730: string theory. Large parts of this section are based on joint work
731: with J.~Polchinski~\cite{BP}.
732:
733: \subsection{A continuous spectrum of $\Lambda$?}
734:
735: The first task is to show that the cosmological constant can take on a
736: sufficiently dense ``discretuum'' of values. In string theory, each
737: line in the spectrum of $\Lambda$ will correspond to a long-lived
738: metastable vacuum.
739:
740: Why look for a discretuum and not a continuum of values? The quick
741: answer is that we can plausibly realize a discretuum in string theory,
742: but not a continuum. In fact, we know of no adjustable parameters on
743: which the cosmological constant depends in a continuous manner---at
744: least if our goal at the same time is metastability~\cite{BP}.
745:
746: But why insist on metastability? I will give a brief argument that we
747: have good reasons to do so. This shows more generally that it would
748: be difficult to realize the ``variable $\Lambda$'' approach with a
749: continuous spectrum.
750:
751: If the continuous parameter is like an integration constant, fixed
752: once and for all, then it will not allow $\Lambda$ to vary between
753: large regions in the universe, so it would have to be tuned by hand.
754: If the parameter can change over time, then the vacuum energy can be
755: lowered by sliding down the spectrum continuously. But this is
756: tantamount to introducing a scalar field potential, and it leads to
757: versions of problems described in Sec.~\ref{sec-denial}: Why, in ten
758: billion years, has $\Lambda$ not relaxed to its lowest possible value?
759: (We cannot assume that this ``ground state'' is the observed value, or
760: zero, since this would beg the question; radiative corrections would
761: immediately destroy such a setup.)
762:
763: It is difficult to see how such a special behavior could be arranged,
764: other than in a theory with many metastable vacua, but this would get
765: us back to the discrete case. Moreover, even with anthropic
766: constraints there is no reason why $\Lambda$ should change as slowly
767: as current bounds indicate. Thus, one would predict a universe with
768: blatantly time-dependent vacuum energy. In the discretuum, on the
769: other hand, the minimum value of the cosmological constant naturally
770: remains fixed for the lifetime of the metastable vacuum, which can
771: easily exceed ten billion years.
772:
773:
774: \subsection{A single four-form field}
775: \label{sec-single}
776:
777: To begin, I will present a very simple model of a discretuum. This
778: model will not work for two reasons: it cannot be realized in string
779: theory, and it produces an empty universe~\cite{BT1,BT2}.
780: Nevertheless, it will be instructive, and it invites a useful analogy
781: with electromagnetism.
782:
783: Recall that the Maxwell field, $F_{ab}$, is derived from a potential,
784: $F_{ab} = \partial_a A_b - \partial_b A_a$. The potential is sourced
785: by a point particle through a term $\int e \mathbf A$ in the action,
786: where the integral is over the worldline of the particle, and e is the
787: charge. Technically, $\mathbf F$ is a two-form (a totally
788: antisymmetric tensor of rank 2), and $\mathbf A$ is a one-form
789: coupling to a one-dimensional worldvolume (the worldline of the
790: electron).
791:
792: The field content of string theory and supergravity is completely
793: determined by the structure of the theory. It includes a four-form
794: field, $F_{abcd}$, which derives from a three-form potential:
795: \begin{equation}
796: F_{abcd} = \partial_{[a} A_{bcd]}~,
797: \end{equation}
798: where square brackets denote total antisymmetrization. This potential
799: naturally couples to a two-dimensional object, a membrane, through a
800: term $\int q \mathbf A$, where the integral runs over the 2+1
801: dimensional membrane worldvolume, and $q$ is the membrane charge.
802:
803: The properties of the four-form field in our 3+1 dimensional world
804: mirror the behavior of Maxwell theory in a 1+1 dimensional system.
805: Consider, for example, an electric field between two capacitor plates.
806: Its field strength is constant both in space and time. Its magnitude
807: depends on how many electrons the negative plate contains; thus it
808: will be an integer multiple of the electron charge: $E = ne$.
809:
810: Its energy density will be one half of the field strength squared:
811: \begin{equation}
812: \rho = \frac{F_{ab}F^{ab}}{2} = \frac{n^2e^2}{2}
813: \end{equation}
814: In order to treat this as a system with only one spatial dimension, I
815: have integrated over the directions transverse to the field lines, so
816: $\rho$ is energy per unit length. The pressure is equal to $-\rho$.
817: The corresponding 1+1 dimensional stress tensor has the form of
818: Eq.~(\ref{eq-lrho}), so the electromagnetic stress tensor acts like
819: vacuum energy in 1+1 dimensions.
820:
821: The same is true for the four-form in our 3+1 dimensional world.
822: First of all, the equation of motion in the absence of sources is
823: $\partial_a(\sqrt{-g} F^{abcd}) = 0$, with solution
824: \begin{equation}
825: F^{abcd} = c \epsilon^{abcd}~,
826: \end{equation}
827: where $\epsilon $ is the unit totally anti-symmetric tensor and $c$ is
828: an arbitrary constant. In string theory, there are ``magnetic''
829: charges (technically, five-branes) dual to the ``electric charges''
830: (the membranes) sourcing the four-form field. Then, by an analogue of
831: Dirac quantization of the electric charge, one can show that $c$ is
832: quantized in integer multiples of the membrane charge, $q$:
833: \begin{equation}
834: c = nq~.
835: \end{equation}
836: Note that the actual value of the four-form field is thus quantized,
837: not only the difference between possible values.
838:
839: The four-form field strength squares to $F_{abcd}F^{abcd} = 24 c^2$,
840: and the stress tensor is proportional to the metric, with
841: \begin{equation}
842: \rho = \frac{1}{2 \times 4!} F_{abcd}F^{abcd} = \frac{n^2 q^2}{2}
843: \end{equation}
844: In summary, the four-form field is non-dynamical, and it contributes
845: $n^2 q^2/2$ to the vacuum energy. It is thus indistinguishable from a
846: contribution to the cosmological constant.
847:
848: Next, let us include non-perturbative quantum effects. The electric
849: field between the plates will be slowly discharged by Schwinger pair
850: creation of field sources. This is a process by which a electron and
851: a positron tunnel out of the vacuum. Since field lines from the
852: plates can now end on these particles, the electric field between the
853: two particles will be lower by one unit [$ne\rightarrow (n-1)e$]. The
854: particles appear at a separation such that the corresponding decrease
855: in field energy compensates for their combined rest mass. They are
856: then subjected to constant acceleration by the electric field until
857: they hit the plates. If the plates are far away, they will move
858: practically at the speed of light by that time.
859:
860: For weak fields, this tunneling process is immensely suppressed, with
861: a rate of order $\exp(-\pi m^2/ne^2)$, where the exponent arises as
862: the action of a Euclidean-time solution describing the appearance of
863: the particles. Thus, a long time passes between creation events.
864: However, over large enough time scales, the electric field will
865: decrease by discrete steps of size $e$. Correspondingly, the 1+1
866: dimensional ``vacuum energy'', i.e., the energy per unit length in the
867: electric field, will decrease by a discrete amount $[n^2
868: e^2-(n-1)^2e^2]/2 = (n-\frac{1}{2})e^2$. Note that this step size
869: depends on the remaining flux.
870:
871: Precisely analogous nonperturbative effects occur for the four-form
872: field in 3+1 dimensions. By an analogue of the Schwinger process,
873: spherical membranes can spontaneously appear. (This is the correct
874: analogue: the two particles above form a zero-sphere, i.e., two
875: points; the membrane forms a two-sphere.) Inside this source, the
876: four-form field strength will be lower by one unit of the membrane
877: charge [$nq\rightarrow (n-1)q$]. The process conserves energy: the
878: initial membrane size is such that the membrane mass is balanced
879: against the decreased energy of the four-form field inside the
880: membrane. The membrane quickly grows to convert more space to the
881: lower energy density, expanding asymptotically at the speed of light.
882:
883: Membrane creation is a well-understood process described by a
884: Euclidean instanton, and like Schwinger pair creation, is generically
885: exponentially slow. Ultimately, however, it will lead to the
886: step-by-step decay of the four-form field. Inside a new membrane, the
887: vacuum energy will be lower by $(n-\frac{1}{2})q^2$.
888:
889: This suggests a mechanism for cancelling off the cosmological
890: constant. Let us collect all contributions (see Sec.~\ref{sec-ccp}),
891: except for the four-form field, in a ``bare'' cosmological constant
892: $\lambda$. Generically, $|\lambda|$ should be of order unity (at
893: least in the absence of supersymmetry), and we will assume without
894: excessive loss of generality that it is negative. With $n$ units of
895: four-form flux turned on, the full cosmological constant will be given
896: by
897: \begin{equation}
898: \Lambda = \lambda+ \frac{1}{2} n^2 q^2
899: \end{equation}
900:
901: If $n$ starts out large, the cosmological constant will decay by
902: repeated membrane creation, until it is close to zero. The smallest
903: value of $|\Lambda|$ is attained for the flux $n_{\rm best}$, given by
904: the nearest integer to $\sqrt{2|\lambda|}/q$. The step size near
905: $\Lambda=0$ is thus given by $(n_{\rm best}-\frac{1}{2})q^2$. For
906: this mechanism to produce a value in the Weinberg window,
907: Eq.~(\ref{eq-weinberg}), this step size would need to be of order
908: $10^{-121}$ or smaller. This requires an extremely small membrane
909: charge,
910: \begin{equation}
911: q< 10^{-121} |\lambda|^{-1/2}
912: \end{equation}
913: (the bare cosmological constant $\lambda$ is at best of order one).
914:
915: This leads to two problems~\cite{BT1,BT2}: the small-charge problem,
916: and the empty-universe problem. The membrane charge $q$ is now itself
917: exceedingly small and thus unnatural. In particular, despite attempts
918: in this direction~\cite{FenMar00}, it is not known how to realize such
919: a small charge in string theory.
920:
921: Assuming the small-charge problem could be resolved, the mechanism
922: would lead to a universe very different from ours: it would be devoid
923: of all matter and radiation. The point is that small values of
924: $\Lambda$ are approached very gradually from above. Thus the universe
925: is dominated by positive vacuum energy all along, leading to
926: accelerated expansion. The exponential suppression of membrane
927: nucleation events ensures that this expansion goes on long enough to
928: dilute all matter. Eventual membrane nucleation decreases the vacuum
929: energy only by a tiny amount ($10^{-121}$ or less). At best, this
930: might reheat the universe to $10^{-30}$, or about
931: $10^{-2}$eV.\footnote{The actual number is vastly smaller still, since
932: most of the energy goes into accelerating the growth of the membrane
933: bubble. This is the reason why the empty-universe problem also
934: plagues ``old inflation''~\cite{GutWei83}, even though the jump in
935: vacuum energy is considerably larger in that case.} This falls well
936: short of the $10$ MeV mark necessary to make contact with standard
937: cosmology, a theory we trust at least back to nucleosynthesis.
938:
939:
940: \subsection{Multiple four-form fields}
941: \label{sec-multiple}
942:
943: The above problems can be overcome by considering a theory with more
944: than one species of four-form field. I will explain why this
945: situation arises naturally in string theory, but first I will discuss
946: how multiple four-form fields can produce a dense discretuum without
947: requiring small charges.
948:
949: Consider a theory with $J$ four-form fields. Correspondingly there
950: will be $J$ types of membrane, with charges $q_1,\ldots,q_J$. Above
951: I analyzed the case of a single four-form field; essentially the
952: conclusions still apply to each field separately. In particular, each
953: field strength separately will be constant in 3+1 dimensions,
954: \begin{equation}
955: F^{abcd}_{(i)} = n_i q_i \epsilon^{abcd}~,
956: \end{equation}
957: and it will contribute like vacuum energy to the stress tensor.
958:
959: Let us again collect all contributions to vacuum energy, {\em
960: except\/} for those from the $J$ four-form fields, in a bare
961: cosmological constant $\lambda$, which I assume to be negative but
962: otherwise generic (i.e., of order unity). Then the total cosmological
963: constant will be given by
964: \begin{equation}
965: \Lambda = \lambda + \frac{1}{2}\sum_{i=1}^J n_i^2 q_i^2~.
966: \label{eq-lmult}
967: \end{equation}
968: This will include a value in the Weinberg window,
969: Eq.~(\ref{eq-weinberg}), if there exists a set of integers $n_i$ such
970: that
971: \begin{equation}
972: 2|\lambda|<\sum n_i^2 q_i^2<2(|\lambda|+ \Delta\Lambda)~,
973: \end{equation}
974: where $\Delta\Lambda\approx 10^{-121}$.
975:
976: A nice way to visualize this problem is to consider a $J$-dimensional
977: grid, with axes corresponding to the field strengths $n_i q_i$, as
978: shown in Fig.~\ref{fig-grid}.
979: %
980: \EPSFIGURE{fig-grid.eps,width=.7\textwidth}{\label{fig-grid} Possible
981: configurations of the four-form fluxes correspond to discrete points
982: in a $J$-dimensional grid. By Eq.~(\ref{eq-lmult}), vacua that allow
983: for structure formation lie within a thin shell of radius
984: $\sqrt{2|\lambda|}$ and width $\Delta\Lambda/\sqrt{2|\lambda|}$, where
985: $\lambda$ is the bare cosmological constant and $\Delta\Lambda$ is
986: the width of the Weinberg window, Eq.~(\ref{eq-weinberg}).}
987: %
988: Every possible configuration of the
989: four-form fields corresponds to a list of integers $n_i$, and thus to
990: a discrete grid point. The Weinberg window can be represented as a
991: thin shell of radius $\sqrt{2|\lambda|}$ and width
992: $\Delta\Lambda/\sqrt{2|\lambda|}$. The shell has volume
993: \begin{equation}
994: V_{\rm shell} = \Omega_{J-1} (\sqrt{2|\lambda|})^{J-1}
995: \frac{\Delta\Lambda}{\sqrt{2|\lambda|}} = \Omega_{J-1}
996: |2\lambda|^{\frac{J}{2}-1}\Delta\Lambda~,
997: \end{equation}
998: where $\Omega_{J-1} = 2\pi^{J/2}/\Gamma(J/2)$ is the area of a unit
999: $J-1$ dimensional sphere. The volume of a grid cell is
1000: \begin{equation}
1001: V_{\rm cell}=\prod_{i=1}^J q_i~.
1002: \end{equation}
1003: There will be at least one value of $\Lambda$ in the Weinberg window,
1004: if $V_{\rm cell}<V_{\rm shell}$, i.e., if
1005: \begin{equation}
1006: \frac{\prod_{i=1}^J q_i}{\Omega_{J-1}
1007: |2\lambda|^{\frac{J}{2}-1}} < |\Delta\Lambda|~.
1008: \label{eq-bp}
1009: \end{equation}
1010:
1011: The most important consequence of this formula is that charges no
1012: longer need to be very small. I will shortly argue that in string
1013: theory one naturally expects $J$ to be in the hundreds. With $J=100$,
1014: for example, Eq.~(\ref{eq-bp}) can be satisfied with charges $q_i$ of
1015: order $10^{-1.6}$, or $\sqrt{q_i}\approx 1/6$ (the latter has mass
1016: dimension 1 and so seems an appropriate variable for the judging
1017: naturalness of this scenario). Interestingly, the large expected
1018: value of the bare cosmological constant is actually welcome: it
1019: becomes more difficult to satisfy Eq.~(\ref{eq-bp}) if $|\lambda|\ll
1020: 1$.
1021:
1022: The origin of the large number of four-form fields lies in the
1023: topological complexity of small extra dimensions. String theory is
1024: most naturally formulated in 9+1 or 10+1 spacetime dimensions. For
1025: definiteness I will work with the latter formulation (also known as
1026: M-theory). If it describes our world, then 7 of the spatial
1027: dimensions must be compactified on a scale that would have eluded our
1028: most careful experiments. Thus one can write the spacetime manifold as
1029: a direct product:
1030: \begin{equation}
1031: M = M_{3+1}\times X_7~.
1032: \end{equation}
1033: Typically, the compact seven-dimensional manifold $X_7$ will have
1034: considerable topological complexity, in the sense of having large
1035: numbers of non-contractable cycles of various dimensions.
1036:
1037: To see what this will mean for the 3+1 dimensional description,
1038: consider a string wrapped around a one-cycle (a ``handle'') in the
1039: extra dimensions. To a macroscopic observer this will appear as a
1040: point particle, since the handle cannot be resolved. Now, recall that
1041: M-theory contains five-branes, the magnetic charges dual to membranes.
1042: Like strings on a handle, five-branes can wrap higher-dimensional
1043: cycles within the compact extra dimensions. A five-brane wrapping a
1044: three-cycle (a kind of non-contractible three-sphere embedded in the
1045: compact manifold) will appear as a two-brane, i.e., a membrane, to the
1046: macroscopic observer.
1047:
1048: Six-dimensional manifolds, such as Calabi-Yau geometries, generically
1049: have hundreds of different three-cycles, and adding another dimension
1050: will only increase this number. The five-brane---one of a small
1051: number of fundamental objects of the theory---can wrap any of these
1052: cycles, giving rise to hundreds of apparently different membrane
1053: species in 3+1 dimensions, and thus, to $J\sim O(100)$ four-form
1054: fields, as required.
1055:
1056: The charge $q_i$ is determined by the five-brane charge (which is set
1057: by the theory to be of order unity), the volume of $X_7$, and the
1058: volume of the $i$-th three-cycle. The latter factors can lead to
1059: charges that are slightly smaller than 1, which is all that is
1060: required. Note also that the volumes of the three-cycles will
1061: generically differ from each other, so one would expect the $q_i$ to
1062: be mutually incommensurate. This is important to avoid huge
1063: degeneracies in Eq.~(\ref{eq-lmult}).
1064:
1065: Each of the flux configurations ($n_1,\ldots,n_J$) corresponds to a
1066: metastable vacuum. Fluxes can only change if a membrane is
1067: spontaneously created. As discussed in Sec.~\ref{sec-single}, this
1068: Schwinger-like process is generically exponentially suppressed,
1069: leading to extremely long lifetimes. Thus, multiple four-forms
1070: naturally give a dense discretuum of metastable vacua.
1071:
1072: The model I have presented is an oversimplification. When it was
1073: first proposed, it was not yet understood how to stabilize the compact
1074: manifold against deformations (technically, how to fix all moduli
1075: fields including the dilaton). This is clearly necessary in any case
1076: if string theory is to describe our world. But one would expect that
1077: in a realistic compactification, the fluxes wrapped on cycles should
1078: deform the compact manifold, much like a rubber band wrapping a
1079: doughnut-shaped balloon. Yet, I have pretended that $X_7$ stays
1080: exactly the same independently of the fluxes $n_i$.
1081:
1082: Therefore, Eq.~(\ref{eq-lmult}) will not be correct in a more
1083: realistic model. The charges $q_i$, and indeed the bare cosmological
1084: constant $|\lambda|$, will themselves depend on the integers $n_i$.
1085: Thus the cosmological constant may vary quite unpredicably. But the
1086: crucial point remains unchanged: the number of vacua, $N$, can be
1087: extremely large, and the discretuum should have a typical spacing
1088: $\Delta\Lambda\approx 1/N$. For example, if there are $500$
1089: three-cycles and each can support up to 9 units of flux, there will be
1090: of order $N=10^{500}$ metastable configurations. If their vacuum
1091: energy is effectively a random variable with at most the Planck value
1092: ($|\Lambda|\lesssim 1$), then there will be $10^{380}$ vacua in the
1093: Weinberg window, Eq.~(\ref{eq-weinberg}).
1094:
1095: In the meantime, there has been significant progress with stabilizing
1096: the compact geometry (e.g., Refs.~\cite{DasRaj99,GidKac01}; see
1097: Refs.~\cite{Sil04,Gra05,DouKac06} for reviews.). In particular,
1098: Kachru, Kallosh, Linde, and Trivedi~\cite{KKLT} have shown that
1099: metastable de~Sitter vacua can be realized in string theory while
1100: fixing all moduli.\footnote{Constructions in non-critical string
1101: theory (i.e., string theory with more than ten spacetime dimensions)
1102: were proposed earlier~\cite{Sil01,MalSil02}.} Their construction
1103: supports the above argument that the number of flux vacua can be
1104: extremely large. More sophisticated counting methods~\cite{DenDou04b}
1105: bear out the quantitative estimates obtained from the simple model I
1106: have presented.
1107:
1108: I will close with two remarks. The need for extra dimensions could be
1109: regarded as an unpleasant aspect of string theory, since it forces us
1110: to worry about why and how they are hidden. Ironically, they are
1111: precisely what has allowed string theory to address the cosmological
1112: constant problem and pass its first observational test.
1113:
1114: One often hears that there are now $10^{500}$ ``string theories'',
1115: suggesting a loss of fundamental simplicity and uniqueness. This is
1116: like saying that there are myriads of standard models because there
1117: are many ways to make a lump of iron. From five standard model
1118: particles, one can construct countless metastable configurations of
1119: atoms, molecules, and condensed matter objects. Similarly, the large
1120: number of vacua in string theory arises by combining a small set of
1121: fundamental ingredients in different ways, {\em in the extra
1122: dimensions}. From this perspective, numbers like $10^{500}$ should
1123: not surprise us.
1124:
1125:
1126: \subsection{Our way home}
1127: \label{sec-home}
1128:
1129: I have argued that string theory contains such a dense spectrum of
1130: metastable vacua that many of them will satisfy the Weinberg
1131: inequality, Eq.~(\ref{eq-weinberg}). But still, they represent only a
1132: very small fraction of the total number of vacua. Hence, there is no
1133: particular reason to assume that the universe would have started out
1134: in one of the relatively rare vacua with small late-time cosmological
1135: constant. Such an assumption would be especially problematic since
1136: the late-time value of the cosmological constant is initially far from
1137: apparent. In our own vacuum, for example, the cosmological constant
1138: is now small but was enormously larger at early times, before
1139: inflation ended and various symmetries were broken.
1140:
1141: Fortunately, it is unnecessary to assume that the universe starts out
1142: in a Weinberg vacuum. I will now show that starting from generic
1143: initial conditions, the universe will grow arbitrarily large. Over
1144: time, it will come to contain enormous regions (``bubbles'' or
1145: ``pockets'') corresponding to each metastable vacuum
1146: (Fig.~\ref{fig-global}). In particular, the Weinberg vacua will be
1147: realized somewhere in this ``multiverse''. It will be seen that these
1148: vacua can be efficiently reheated, so the empty-universe problem of
1149: Sec.~\ref{sec-single} will not arise.
1150: %
1151: \EPSFIGURE{fig-global.eps,width=.8\textwidth}{\label{fig-global}
1152: Bird's eye view of the universe. There are regions corresponding to
1153: every vacuum in the landscape (shown in different colors). Each
1154: region is an infinite, spatially open universe; the dashed line
1155: shows an example of an instant of time. The black diamond is an
1156: example of a spacetime region that is causally accessible to a
1157: single observer (see Sec.~\ref{sec-predict}).}
1158: %
1159:
1160: By Eq.~(\ref{eq-lmult}), all but a finite number of metastable vacua
1161: will have $\Lambda>0$. Let us assume that the universe begins in one
1162: of these vacua. Of course, this means that typically the cosmological
1163: constant will be large initially. Since $\Lambda>0$, the universe
1164: will be well described by de~Sitter space. It can be thought of as a
1165: homogeneous, isotropic universe expanding exponentially on a
1166: characteristic time scale $\Lambda^{-1/2}$.
1167:
1168: Every once in a long while (this time scale being set by the action of
1169: a membrane instanton, and thus typically much larger than
1170: $\Lambda^{-1/2}$), a membrane will spontaneously appear and the
1171: cosmological constant will jump by $(n_i-\frac{1}{2}) q_i^2$. But
1172: this does not affect the whole universe. $\Lambda$ will have changed
1173: only inside the membrane bubble. This region grows arbitrarily large
1174: as the membrane expands at the speed of light.
1175:
1176: But crucially, this does {\em not}\/ imply that the whole universe is
1177: converted into the new vacuum~\cite{ColDel80}. This technical result
1178: can be understood intuitively. The ambient, old vacuum is still, in a
1179: sense, expanding exponentially fast. The new bubble eats up the old
1180: vacuum as fast as possible, at nearly the speed of light. But this is
1181: not fast enough to compete with the background expansion.
1182:
1183: More and more membranes, of up to $J$ different types, will nucleate
1184: in different places in the rapidly expanding old vacuum. Yet, there
1185: will always be some of the old vacuum left. One can show that the
1186: bubbles do not ``percolate'', i.e., they will never eat up all of
1187: space~\cite{GutWei83}. Thus different fluxes can change, and
1188: different directions in the $J$-dimensional flux space are explored.
1189:
1190: Inside the new bubbles, the game continues. As long as $\Lambda$ is
1191: still positive, there is room for everyone, because the background
1192: expands exponentially fast. In this way, all the points in the flux
1193: grid $(n_1,\ldots,n_J)$, are realized as actual regions in physical
1194: space. The cascade comes to an end wherever a bubble is formed with
1195: $\Lambda<0$, but this affects only the interior of that particular
1196: bubble (it will undergo a big crunch). Globally, the cascade
1197: continues endlessly.
1198:
1199: Perhaps surprisingly, each bubble interior is an open FRW universe in
1200: its own right, and thus infinite in spatial extent.\footnote{In an
1201: open universe, spatial hypersurfaces of constant energy density are
1202: three-dimensional hyperboloids. This shape is dictated by the
1203: symmetries of the instanton describing the membrane nucleation. It
1204: is closely related to the hyperbolic shape of the spacetime paths of
1205: accelerating particles, like the electron-positron pair studied
1206: above.} Yet, each bubble is embedded in a bigger universe (sometimes
1207: called ``multiverse'' or ``megaverse''), which is extremely
1208: inhomogeneous on the largest scales.
1209:
1210: An important difference to the model with only one four-form is that
1211: the vacua will not be populated in the order of their vacuum energy.
1212: Two neighboring vacua in flux space (i.e., neighbors in the
1213: ``landscape''), will differ hugely in cosmological constant. That is,
1214: they differ by one unit of flux, and the charges $q_i$ are not much
1215: smaller than one, so by Eq.~(\ref{eq-lmult}) this translates into an
1216: enormous difference in cosmological constant. Conversely, vacua with
1217: very similar values of the cosmological constant will be well
1218: separated in the flux grid (i.e., far apart in the landscape).
1219:
1220: This feature is crucial for solving the empty universe problem. When
1221: our vacuum was produced in the interior of a new membrane, the
1222: cosmological constant may have decreased by as much as $1/100$ of the
1223: Planck density. Hence, the temperature before the jump was enormous
1224: (in this example, the Gibbons-Hawking temperature of the corresponding
1225: de~Sitter universe would have been of order $1/10$ of the Planck
1226: temperature), and only extremely massive fields will have relaxed to
1227: their minima. Most fields will be thermally distributed and can only
1228: begin to approach equilibrium after the jump decreases the vacuum
1229: energy to near zero.
1230:
1231: Thus, the final jump takes on the role analogous to the big bang in
1232: standard cosmology. The ``universe'' (really, just our particular
1233: bubble) starts out hot and dense. If the effective theory in the
1234: bubble contains scalar fields with suitable potentials, there will be
1235: a period of slow-roll inflation as their vacuum energy slowly relaxes.
1236: (This was apparently the case in our vacuum.) At the end of this
1237: slow-roll inflation process, the universe reheats.
1238:
1239: To a (purely hypothetical) observer in the primordial era of a given
1240: bubble, it would be far from obvious what the late-time cosmological
1241: constant will be, since this depends on future symmetry breakings and
1242: the relaxation of scalar field potential energy. The small late-time
1243: values in some bubbles are the result of purely accidental
1244: cancellations---which are bound to happen in some vacua if there are
1245: $10^{500}$ vacua in total.
1246:
1247: To a hypothetical primordial observer in our own bubble, the evolution
1248: of vacuum energy would seem like a sequence of bizzare coincidences.
1249: I assume here that the observer is sufficiently intelligent to know
1250: that quantum field theory predicts a cosmological constant of order
1251: one. In the primordial era, the energy density in radiation is large,
1252: and it could mask even a fairly large cosmological constant. But as
1253: the universe cools off, a cosmological constant exceeding the ever
1254: decreasing energy density in matter and radiation would become
1255: immediately apparent. Thus, the discrepancy between theory and
1256: observation grows larger and larger.
1257:
1258: Much to his surprise, our observer would find the vacuum energy in the
1259: minimum of the inflationary potential to be much smaller than during
1260: inflation---in fact, it cannot be distinguished from zero. (This
1261: allows the universe to reheat, without immediately inflating all
1262: matter away, but why would our observer care?) During electroweak
1263: symmetry breaking, at time $10^{-12}$ sec, the vacuum energy density
1264: shifts by $(200$ GeV$)^4$. Our observer computes this and is thus led
1265: to expect that soon afterwards, when the radiation energy drops below
1266: $(200$ GeV$)^4$, the dynamical effects of a cosmological constant will
1267: finally become apparent. It does not, so the observer is forced to
1268: conclude that the shift must have cancelled against another, equally
1269: large contribution that he had not noticed earlier since radiation was
1270: too dense. In fact, the cancellation is so exquisite that vacuum
1271: energy remains dynamically irrelevant at the much later time $1$ sec.
1272: (This allows nucleosynthesis to proceed.) After hundreds of millions
1273: of years, at vastly lower energy density, still no vacuum energy is
1274: apparent (allowing for the formation of galaxies to proceed
1275: undisturbed). Only after billions of years (after structure has
1276: formed), does vacuum energy resurface and begin to dominate over the
1277: ever more dilute matter energy density.
1278:
1279: If such hypothetical observers existed, this sequence really {\em
1280: would\/} be bizzare and unexpected. There are far more vacua with
1281: similar primordial evolution but without the anomalously small
1282: late-time cosmological constant. All the corresponding bubbles would
1283: presumably harbor similar primordial observers. Then the vast
1284: majority of observers would {\em not\/} see a sequence of ``miracles''
1285: leading to a late-time cosmological constant as small as $10^{-121}$.
1286:
1287: But it appears that no such hypothetical primordial observers exist.
1288: Observers will arise only after some structure has formed. This
1289: happens only in the ``bizzare'', rare vacua in which accidental
1290: cancellations produce a late-time cosmological constant of order
1291: $10^{-121}$ or less. Any larger, and vacuum energy would disrupt
1292: galaxy formation. We should not be surprised, therefore, to find
1293: ourselves in such a bubble.
1294:
1295:
1296:
1297: \section{The landscape and predictivity}
1298: \label{sec-predict}
1299:
1300: \subsection{A new challenge}
1301:
1302: A good explanation will do more than solve a problem. It should offer
1303: us a new way of thinking, and in doing so, raise new, interesting
1304: problems. In fact, the picture I have outlined does present a
1305: tremendous challenge: how does one make predictions in the landscape?
1306:
1307: Let us suppose that there are $10^{500}$ metastable vacua. Among
1308: them, everything varies: forces, coupling strengths, masses, field
1309: content, gauge groups, and other aspects of the low energy
1310: Lagrangian. Are the ``constants'' of nature we measure
1311: constrained by nothing but the fact of our existence? This would be a
1312: bleak prospect indeed.
1313:
1314: In order to look at the problem dispassionately, it helps to take
1315: recourse once more to the analogy with complex, many-particle systems
1316: developed near the end of the previous section. A vast number of
1317: phenomena arise from a few particles in the standard model: the world
1318: is a rich, complex place. But this does not imply that anything goes.
1319: There are only a finite number of elements, and a random combination
1320: of atoms is unlikely to form a stable molecule. Even quantities such
1321: as material properties ultimately derive from standard model
1322: parameters and cannot be arbitrarily dialed.
1323:
1324: Similarly, one would expect that there are low-energy Lagrangians that
1325: simply cannot arise from string theory with its limited set of
1326: ingredients, no matter how complicated the manner in which they are
1327: combined~\cite{Vaf05,ArkMot06,OogVaf06}.
1328:
1329: Moreover, the great complexity of a system need not be an obstacle to
1330: its effective description. Imagine we had never heard of
1331: thermodynamics and were told to describe the behavior of all the air
1332: molecules in a room. Or suppose we were ignorant of condensed matter
1333: physics, and were charged with deriving the properties of metals from
1334: the standard model. Would we not worry, for a moment, that these
1335: tasks are too complex to be tractable? Of course, we know well that
1336: such problems yield to the laws of large numbers. The predictive
1337: power of statistical or effective theories is completely deterministic
1338: in practice: not in ten billion years will the air ever collect in one
1339: corner of the room. This is not to say that finding such descriptions
1340: is trivial, only that it is possible.
1341:
1342: Similarly, there is every reason to hope that a set of $10^{500}$
1343: vacua will yield to statistical reasoning, allowing us to extract
1344: predictions. Yet we must not presume this task simple or even
1345: straightforward. We are just beginning, so the present scarcity of
1346: predictions is hardly proof of their impossibility.
1347:
1348: The problem can be divided into three separate tasks:
1349: \begin{enumerate}
1350: \item{Statistical properties of the string theory landscape}
1351: \item{Selection effects from cosmological dynamics}
1352: \item{Anthropic selection effects}
1353: \end{enumerate}
1354: The first of these has been tackled by a number of authors; see, e.g.,
1355: Refs.~\cite{Dou03,DenDou04b,GmeBlu05,DouTay06}, or Ref.~\cite{Kum06}
1356: for a review. The question is, what is the relative abundance of
1357: stable or metastable vacua with specified low-energy properties. Our
1358: understanding of metastable vacua is still rather qualitative, so most
1359: investigations focus on supersymmetric vacua instead, which are under
1360: far better control. Clearly, it would be desirable to extend our
1361: samples; this will likely require significant progress in
1362: understanding vacua without supersymmetry. Meanwhile it will be
1363: interesting to understand the extent to which current samples are
1364: representative of more realistic vacua, especially since one is
1365: usually working in a particular corner of moduli space.
1366:
1367: This remains a very active area of research, and I will not attempt a
1368: more detailed review. Next, I will discuss a recent approach to the
1369: second and third task.
1370:
1371:
1372: \subsection{Probabilities in eternal inflation}
1373:
1374: It is not enough to calculate the probability that a random metastable
1375: vacuum picked from the theory landscape has a given property.
1376: Cosmological dynamics is interposed between the theory landscape and
1377: the actual realization of vacua as large regions in the universe.
1378: This dynamical process may preferentially produce some vacua and
1379: suppress others. This is the second question listed above: What is
1380: the relative abundance of different vacua {\em in the physical
1381: universe\/}?
1382:
1383: Computational difficulties aside, this question turns out to be hard
1384: to answer even in principle, because of a scourge of infinities. The
1385: global structure of the universe arising from the string landscape is
1386: extremely complicated (see Sec.~\ref{sec-home}). Each vacuum $i$ is
1387: realized infinitely many times as a bubble embedded in the global
1388: spacetime. Moreover, every bubble is an open universe and thus of
1389: infinite spatial extent.
1390:
1391: The most straightforward way of regulating the infinities is to
1392: consider the universe at finite time before taking a limit. There is
1393: an ambiguity in whether one should compare the volumes, or simply the
1394: number of each type of bubble on this time slice (or some intermediate
1395: quantity). Worse, results depend strongly on the choice of time
1396: variable~\cite{LinLin94,GarLin94}, and no preferred time-slicing is
1397: available in the highly inhomogeneous global spacetime.
1398:
1399: A number of slicing-invariant probability measures have been proposed;
1400: see, e.g.~\cite{GarVil01,GarSch05,EasLim05} for recent work. Yet,
1401: slicing invariance is far from a strong enough criterion for
1402: determining a unique measure; for example, any function of an
1403: invariant measure will again be invariant.
1404:
1405: In addition to these severe ambiguities, known slicing-invariant
1406: proposals appear to lead to predictions that disagree with
1407: observation~\cite{FelHal05,GarVil05,Pag06,BouFre06b}. The first
1408: problem arises in proposals where the probability carried by a vacuum
1409: is proportional to the factor by which inflation increases the volume.
1410: (This refers to the ordinary slow-roll inflation of
1411: Sec.~\ref{sec-inflation}, not the false-vacuum driven eternal
1412: inflation of Sec.~\ref{sec-home}.) This factor is exponential in the
1413: duration of inflation. In Ref.~\cite{FelHal05} it was argued that
1414: generically, both the number of $e$-foldings and the density
1415: perturbations produced will depend monotonically on parameters of the
1416: inflationary model. Thus, the great weight carried by long periods of
1417: inflation should push the density contrast $\delta\rho/\rho$ towards 0
1418: or 1. One can argue that life would be impossible in a universe with
1419: $\delta\rho/\rho$ too small or too large~\cite{TegRee97}. But
1420: anthropic arguments cannot resolve the paradox. The exponential
1421: preference for extreme values means that we should live dangerously, a
1422: lucky fluctuation in an inhospitable universe. Instead, the density
1423: contrast in our universe appears to be comfortably within the
1424: anthropic window.
1425:
1426: A more severe problem arises, e.g., in the proposal by Garriga et
1427: al.~\cite{GarSch05}: One can show that the overwhelming majority of
1428: observers are not like us but arise from random
1429: fluctuations~\cite{Pag06}. Assuming that we are typical observers (as
1430: we must if we want to make any predictions), this conflicts with
1431: observation. It could be avoided if all vacua that can harbor
1432: observers decay on a timescale not much longer than $\Lambda^{-1/2}$.
1433: But this is extremely implausible in the string
1434: landscape~\cite{BouFre06b}.
1435:
1436: Recently, a local (or ``causal'', or ``holographic'') approach has
1437: been developed which avoids the ambiguities and resolves the paradoxes
1438: described above~\cite{Bou06,BouFre06,BouFre06b,BouHar06}. Its
1439: original motivation, however, comes from the study of black hole
1440: evaporation, which appears to be a unitary
1441: process~\cite{StrVaf96,Mal97}. A different kind of paradox arose in
1442: this context: The initial quantum state is duplicated, appearing at
1443: the same instant of time both in the Hawking radiation and inside the
1444: black hole. However, causality prevents any observer from seeing both
1445: copies. Thus, the black hole paradox is resolved if we give up on
1446: trying to describe the spacetime globally~\cite{SusTho93,Pre92}.
1447: Indeed, all that is needed is a theory that can describe the
1448: experience of any observer (as opposed to a theory describing
1449: correlations between points remaining forever out of causal contact,
1450: making predictions which cannot be verified even in principle). But
1451: if the global point of view must be rejected in the context of black
1452: holes, why should it be retained in cosmology?
1453:
1454: From a local point of view, eternal inflation looks quite
1455: different~\cite{BouFre06}. Let us attempt to describe only a single
1456: (though arbitrary) causally connected region. This can be defined as
1457: a ``causal diamond'': the overlap between the causal future and the
1458: causal past of a worldline~\cite{Bou00a}. As seen in
1459: Fig.~\ref{fig-global}, this restriction eliminates most of the global
1460: spacetime. In particular, eternal inflation is no longer eternal.
1461:
1462: Consider a geodesic worldline, starting in some initial vacuum $o$
1463: with large positive cosmological constant. (Really, I am considering
1464: an ensemble of worldlines and regions causally connected to them, in
1465: the sense usually adopted to give meaning to probabilities in quantum
1466: mechanics: identical copies of a system. I am {\em not\/} demanding
1467: that the members of this ensemble coordinate their evolution so as to
1468: fit together and form a well-defined global spacetime.) Since the
1469: probability to do so is nonzero, the worldline eventually enters a
1470: vacuum of zero or negative cosmological constant, from which it will
1471: decay no further.\footnote{If $\Lambda$ vanishes exactly then the
1472: vacuum is presumably supersymmetric and stable. If $\Lambda<0$ the
1473: open universe collapses in a big crunch after a time of order
1474: $\Lambda^{-1/2}$, which is likely to be faster than any further
1475: decay channels.} But which vacua the worldline passes through, on
1476: its way to a ``terminal'' vacuum, is a matter of probability.
1477:
1478: The probability for the worldline to enter vacuum $i$, $p_i$, is
1479: proportional to the expected number of times it will enter vacuum $i$.
1480: This can be computed straightforwardly, and unambiguously, from the
1481: matrix of transition rates between vacua~\cite{Bou06}.
1482:
1483: The probabilities $p_i$ depend on the initial probability distribution
1484: for the vacuum in which the worldline starts out, as one would expect
1485: in most dynamical systems. Inflation does not remove the need for a
1486: theory of initial conditions. I will not address this question here,
1487: except to say that I find it plausible that the universe began in a
1488: vacuum with large cosmological constant, and was equally likely to
1489: start in any such vacuum. The vast majority of vacua will have large
1490: cosmological constant, so this is not a strong assumption.
1491:
1492: The resulting probability measure is predictive. In the semiclassical
1493: regime, decays tend to be exponentially suppressed, so that one decay
1494: channel typically dominates completely in any given vacuum. One would
1495: expect that a number of decays have to happen before the worldline
1496: enters a vacuum on the Weinberg shell, and that the fast decays happen
1497: first. For example, in a model of the type described in
1498: Sec.~\ref{sec-multiple}, the production of a membrane of type $i$ is
1499: less suppressed if the background has more than one unit of the
1500: corresponding flux ($n_i>1$), or if the charge $q_i$ associated with
1501: the membrane is relatively small. One thus predicts that the number
1502: of units of flux should be $0$ or $1$ for most fluxes in our vacuum,
1503: and that we are unlikely to find fluxes associated to small charges
1504: turned on~\cite{BouYan06}.
1505:
1506: The paradox of Ref.~\cite{FelHal05} is resolved because the size of
1507: the causal diamond is cut off by the cosmological constant. It will
1508: never become larger than the horizon in a given vacuum, no matter how
1509: much slow-roll inflation occurs after the corresponding bubble is
1510: formed. Thus, exponentially large expansion factors do not enter.
1511: This does not mean that the measure is insensitive to the important
1512: question of whether inflation occurs. However, that issue arises only
1513: if we ask about the suitability of vacua for observers. I will turn
1514: to this question next.
1515:
1516:
1517: \subsection{Beyond the anthropic principle}
1518:
1519: Most vacua will not contain observers. This statement is not
1520: particularly controversial: for example, most vacua will have a
1521: cosmological constant of order unity, and hence will not give rise to
1522: causally connected regions much larger than a Planck length. Entropy
1523: bounds~\cite{CEB1,CEB2} imply that such regions contain at most a few
1524: degrees of freedom, and only a few bits of information. This rules
1525: out complex structures.
1526:
1527: Therefore, the probability for a worldline to enter a given vacuum,
1528: $p_i$, is not the same thing as the probability for that vacuum to be
1529: observed, $\pi_i$. Let us define a weight $w_i$ that measures (in a
1530: sense to be quantified below) the chance that the vacuum $i$ contains
1531: observers. Then
1532: \begin{equation}
1533: \pi_i = \frac{p_i w_i}{\sum p_j w_j}~.
1534: \end{equation}
1535:
1536: Estimating the weights $w_i$ is awkward for a number of reasons. The
1537: biggest difficulty is to define what we mean by an ``observer''. And
1538: given a definition, it can still be extremely hard to estimate whether
1539: observers will form in a given vacuum. What we can do reasonably well
1540: is to consider hypothetical, small changes of one or two of the
1541: parameters describing our own vacuum, and compute their effect on the
1542: formation of life like ours. But this is of little use for estimating
1543: the weights $w_i$ of other vacua in the landscape, since they
1544: generically have radically different low-energy physics. Some
1545: correlations may appear quite robust, such as Weinberg's assertion
1546: that some kind of structure formation is a prerequisite for observers.
1547: But others seem hopelessly specific. For example, can we seriously
1548: expect that life requires carbon? What would this statement even mean
1549: in a low-energy theory with a different standard model gauge group?
1550:
1551: In the global approach, an additional difficulty arises: Strictly,
1552: $w_i$ is either $0$ or $1$. Either there are observers in vacuum $i$,
1553: or there are not. Intuitively, this seems too crude; there should be
1554: a more nuanced sense in which some vacua can be more or less
1555: hospitable to life. But how would we tell whether a vacuum contains
1556: more observers than another? Each bubble is an infinite homogeneous
1557: open universe. At all times, the spatial volume is strictly infinite.
1558: So if observers can form at all, there will be an infinite number of
1559: them. (A method for dealing with this problem within the global
1560: apprach has been suggested in Ref.~\cite{GarVil01}.)
1561:
1562: In the local approach, this problem does not arise. The causal
1563: diamond will be at most of linear size $|\Lambda_i|^{-1/2}$, where
1564: $\Lambda_i$ is the cosmological constant in vacuum $i$. (I will
1565: ignore vacua with vanishing cosmological constant, since they would
1566: have to be exactly supersymmetric, ruling them out as hosts of complex
1567: structures.) Thus, the causally connected region is automatically
1568: finite, providing a natural cutoff.
1569:
1570: The local approach can also help overcome the problem of the excessive
1571: specificity of anthropic considerations~\cite{Bou06}. The key idea is
1572: that observers, whatever they may consist of, need to be able to
1573: increase the entropy. It is implausible that complex systems like
1574: observers will still operate when everything has thermalized and all
1575: free energy has been used up. Everything interesting happens while
1576: the universe returns to equilibrium after the phase transition
1577: associated with the formation of a new bubble.\footnote{This is the
1578: reason why I defined the $p_i$ to be the probability for the
1579: worldline to {\em enter\/} vacuum $i$, rather than the expected
1580: amount of time the worldline will spend in vacuum $i$. The latter
1581: will typically be exponentially greater than the thermalization time
1582: scale and hence is of no relevance.}
1583:
1584: Let us assume that every binary operation will increase the entropy by
1585: at least an amount of order unity~\cite{KraSta00}. On average, one
1586: would expect the number of observers to be related to the total amount
1587: by which entropy increases in a given vacuum. Of course, in the
1588: global viewpoint this statement would be nonsense: if the entropy
1589: increases at all, it will increase by an infinite amount over the
1590: infinite open space. In the local viewpoint, the entropy increase is
1591: not only finite but can be very sharply defined in terms of the causal
1592: diamonds themselves.
1593:
1594: The entropy increase is the difference between the entropy entering
1595: the diamond through the bottom cone, $S_{\rm in}$, and the entropy
1596: leaving through the top cone, $S_{\rm out}$, as shown in
1597: Fig.~\ref{fig-deltas}:
1598: \begin{equation}
1599: \Delta S = S_{\rm out}-S_{\rm in}~.
1600: \end{equation}
1601: The proposal is to weight each vacuum by the entropy increase it admits
1602: \begin{equation}
1603: w_i = \Delta S(i)~.
1604: \end{equation}
1605: Two observers will increase the entropy twice as much as one, so I
1606: have chosen a linear weighting. (There may be nonlinear effects, for
1607: example a sharp cutoff on the minimum entropy increase required to
1608: have at least one observer; smaller $\Delta S$ would be assigned
1609: weight zero.)
1610: %
1611: \EPSFIGURE{fig-deltas.eps,width=.5\textwidth}{\label{fig-deltas}
1612: Instead of explicit anthropic requirements, a new proposal is to
1613: weight each vacuum by the amount of entropy, $\Delta S$, produced
1614: after reheating. This is the difference between the entropy
1615: entering the bottom cone of the causal diamond, $S_{\rm in}$, and
1616: the entropy going out through the top cone, $S_{\rm out}$.}
1617: %
1618:
1619: To be precise, let us take the tip of the bottom cone to lie on the
1620: reheating surface (if there is one; otherwise, no entropy is produced
1621: in any case). Before this time, the universe is empty, because bubble
1622: formation is strongly suppressed (Sec.~\ref{sec-multiple}). Only
1623: after reheating will there be matter, and it can organize itself no
1624: faster than at the speed of light. The tip of the top cone can be
1625: taken to be at a very late time, or even after the vacuum decays; in
1626: the late-time limit the entropy $S_{\rm out}$ will converge quickly.
1627: In a vacuum with positive cosmological constant, the top cone will
1628: coincide with the de~Sitter horizon.
1629:
1630: I will not include entropy associated with event horizons. This would
1631: dominate, particularly through the contribution from the cosmological
1632: horizon in de~Sitter space. Unlike matter entropy production, it is
1633: not clear how an increase in Bekenstein-Hawking entropy can be related
1634: to the physical process of observation. However, one could consider
1635: including this contribution for formal simplicity. In this case, the
1636: argument for prior-based predictions below would need to be augmented
1637: by the extra assumption that in {\em our\/} vacuum, the entropy
1638: produced when black holes are formed, or when they evaporate, is not
1639: related to observers. The prior-free prediction of $-\log\Lambda$
1640: below would be strengthened, on the other hand, since the horizon
1641: entropy is inversely proportional to $\Lambda$.
1642:
1643: Let us compare this ``entropic'' weighting to the anthropic principle.
1644: The latter has been used to predict quantities (such as the
1645: cosmological constant) based on other parameters of our particular
1646: vacuum (such as the time of galaxy formation). In fact it has {\em
1647: only\/} been used to make such ``prior-based'' predictions. Other
1648: examples (some of which happened to be post-dictions) include bounds
1649: on the density contrast $\delta\rho/\rho$~\cite{TegRee97} and on
1650: curvature~\cite{VilWin96,FreKle05}.
1651:
1652: In this relatively modest arena, the entropic weighting competes very
1653: well~\cite{BouHar06}. It turns out that the entropy increase of our
1654: own vacuum is dominated by the photons produced by stars, giving
1655: $\Delta S\approx 10^{85}$. This means that any variation of
1656: parameters that interferes with star formation will cause $\Delta S$
1657: to drop drastically. For example, if $\Lambda$ were much larger, no
1658: structure would form, and hence no stars would form, so this
1659: possibility is suppressed by a large drop in $\Delta S$. In this way,
1660: the entropic weighting reproduces the successes of the anthropic
1661: principle in bounding $\Lambda$, $\delta\rho/\rho$, and curvature in
1662: terms of observed priors.
1663:
1664: This success is remarkable. The assumptions going into anthropic
1665: arguments are quite specific and detailed. By contrast, the entropic
1666: weighting is based on a single, simple thermodynamic condition that
1667: observers must satisfy: they must be able to increase the entropy.
1668:
1669: In some cases, the entropic weighting will even lead to better
1670: quantitative agreement between predictions and data. Anthropic
1671: arguments still expect the cosmological constant to be about 100 times
1672: larger than observed~\cite{MarSha97,Vil04,Wei05}. Large values of
1673: $\Lambda$ are preferred because there are more such vacua, and the
1674: anthropic cutoff is somewhat above the observed value. In the
1675: entropic weighting, the preference for large $\Lambda$ is weaker: The
1676: overall mass included in the causal diamond scales like
1677: $\Lambda^{-1/2}$. This shifts the preferred value to smaller
1678: $\Lambda$, in better agreement with observation.
1679:
1680: Entropic weighting may allow us to attempt predictions {\em without\/}
1681: priors, a feat thoroughly beyond the ambition of anthropic reasoning.
1682: For example, one might ask where a scale like $10^{-123}$ ultimately
1683: comes from~\cite{Pol06}. Anthropic arguments only relate the
1684: cosmological constant in our vacuum to the time of galaxy formation
1685: our vacuum. But in some other vacuum, perhaps stars could have formed
1686: much earlier, allowing the cosmological constant to be much
1687: larger~\cite{Agu01}.
1688:
1689: In fact, this is a serious concern. In the string landscape, many
1690: parameters vary, including $\Lambda$, but also $\delta\rho/\rho$, the
1691: baryon-to-photon ratio, etc. Taking this into account, is the small
1692: observed value of $\Lambda$ not terribly unlikely after all? Weinberg
1693: showed only that $\Lambda$ could not be much larger {\em if all other
1694: parameters are held fixed}. But they are not, and this may spoil
1695: his explanation of the smallness of the cosmological constant. (It
1696: cannot spoil his prediction, which, quite sensibly, took observed data
1697: into account. But it could shift the mystery to questions such as why
1698: $\delta\rho/\rho$ or the baryon-to-photon ratio are so small.)
1699:
1700: To address this issue, let us define a weight that depends only on
1701: $\Lambda$, with individual vacua "integrated out":
1702: \begin{equation}
1703: W(\Lambda) d\Lambda= \sum w_i = \sum \Delta S(i)~,
1704: \end{equation}
1705: where $i$ runs over all the vacua with cosmological constant between
1706: $\Lambda$ and $\Lambda+d\Lambda$. Here $d\Lambda$ should be chosen
1707: large enough for the sum to include a large number of vacua. Thus
1708: $W(\Lambda)$ is an average weight as a function of $\Lambda$.
1709:
1710: The individual weights in this sum will vary hugely. In fact, I would
1711: expect that $\Delta S(i)$ will typically be quite small. That is, it
1712: should be atypical to get inflation and reheating, let alone to
1713: dynamically develop complex processes that produce a lot of entropy
1714: after reheating. But we are interested only in the average of the
1715: weights $w_i$ when summing over a lot of vacua, and in fact we only care
1716: how this average depends on $\Lambda$.
1717:
1718: Let us now make an assumption: suppose that the average is
1719: proportional to (though perhaps much smaller than) the maximum weight
1720: a vacuum can theoretically have, given $\Lambda$. The entropy
1721: difference cannot be greater than the entropy $S_{\rm out}$. This is
1722: turn is bounded by the second law of thermodynamics: it must not
1723: exceed the entropy of the cosmological horizon, which is
1724: $3\pi/\Lambda$. (I am not counting horizon entropy towards $\Delta S$,
1725: since it seems unrelated to the probability of observers, but it can
1726: still be used to bound the entropy produced by matter.)
1727:
1728: In fact this bound can be saturated: the total mass inside the horizon
1729: can be up to $\Lambda^{-1/2}$, and the lowest energy quanta one can burn
1730: it into have wavelength $\Lambda^{-1/2}$, so one can produce up to
1731: $1/\Lambda$ quanta. Of course, one would not expect this extreme limit
1732: to be attained in any significant fraction of vacua (in our own we are
1733: down by $10^{-38}$). The idea is just that the average weight should
1734: scale in the same way with $\Lambda$ as the maximum weight. So this
1735: gives
1736: \begin{equation}
1737: W(\Lambda)\propto \Lambda^{-1}~.
1738: \end{equation}
1739:
1740: Neglecting for a moment the finiteness of the discretuum density, the
1741: probability for $\Lambda$ to be between $a$ and $b$ will thus be
1742: proportional to $\log a - \log b$.
1743:
1744: Now let us assume a discretuum of vacua with roughly even spacing
1745: $1/N$, and $N\approx 10^{500}$. Thus $\log\Lambda$ will range from
1746: $-500$ to $0$. According to the above probability, observers should
1747: find themselves at some generic place in this interval, i.e.,
1748: $-\log\Lambda$ should be $O(100)$.
1749:
1750: Clearly, the assumptions going into this argument warrant further
1751: investigation. Moreover, the result is far less precise than the
1752: Weinberg prediction. This was to be expected when all recourse to
1753: previously measured quantitites is abandoned. But it is reassuring
1754: that quite conceivably, the observed value of the cosmological
1755: constant does not become enormously unlikely, even if all other
1756: parameters are allowed to scan; in fact it remains quite typical.
1757:
1758: More generally, the argument illustrates that even in the landscape,
1759: we need not give up on predicting observable parameters from the
1760: fundamental theory. Under the stated assumptions, the order of
1761: magnitude of the logarithm of the size of the universe is related to
1762: the topological complexity of six-dimensional compact manifolds. This
1763: result is prior free in the sense that it does not use properties of
1764: any particular vacuum, just the structure of the theory.
1765:
1766: \acknowledgments I would like to thank many colleagues for useful
1767: discussions, especially A.~Aguirre, B.~Freivogel, L.~Hall, R.~Harnik,
1768: G.~Kribs, A.~Linde, G.~Perez, J.~Polchinski, M.~Porrati, and I.~Yang.
1769:
1770:
1771: \bibliographystyle{board}
1772: \bibliography{all}
1773: \end{document}
1774: