0712.3324/p.tex
1: \documentclass[prd,twocolumn,nofootinbib]{revtex4}
2: 
3: \usepackage{amsmath,amssymb,graphicx,subfigure}
4: 
5: \begin{document}
6: 
7: \preprint{\hepth{0712.3324}}
8: 
9: \title{Boltzmann babies in the proper time measure}
10: 
11: \author{Raphael Bousso, Ben Freivogel and I-Sheng Yang\footnote{bousso@lbl.gov, freivogel@berkeley.edu, 
12: jingking@berkeley.edu}}
13: 
14: \affiliation{Department of Physics and Center for Theoretical
15: Physics \\
16: University of California, Berkeley, CA 94720, U.S.A. \\
17: {\em and}\\
18: Lawrence Berkeley National Laboratory, Berkeley, CA 94720, U.S.A. }
19: 
20: \begin{abstract}
21:   After commenting briefly on the role of the typicality
22:   assumption in science, we advocate a phenomenological approach to
23:   the cosmological measure problem.  Like any other theory, a measure
24:   should be simple, general, well-defined, and consistent with
25:   observation.  This allows us to proceed by elimination.  As an
26:   example, we consider the proper time cutoff on a geodesic
27:   congruence.  It predicts that typical observers are quantum
28:   fluctuations in the early universe, or Boltzmann babies.  We sharpen
29:   this well-known youngness problem by taking into account the
30:   expansion and open spatial geometry of pocket universes.  Moreover,
31:   we relate the youngness problem directly to the probability
32:   distribution for observables, such as the temperature of the cosmic
33:   background radiation.  We consider a number of modifications of the
34:   proper time measure, but find none that would make it compatible
35:   with observation.
36: \end{abstract}
37: 
38: \maketitle
39: 
40: \section{Introduction}
41: 
42: \subsection{Typicality}
43: 
44: Every time we interpret an experiment, we assume that we are a typical
45: observer.  Suppose, for example, that we are trying to distinguish
46: between two theories $T_1$ and $T_2$.  Conveniently, they predict a
47: very different value of the spin of an electron subjected to a
48: suitable sequence of interactions: $T_1$ predicts spin up with
49: probability $\epsilon$, and $T_2$ predicts spin down with probability
50: $\epsilon$.  
51: 
52: If $\epsilon\ll 1$, then even a single measurement will allow us to
53: rule out one of these theories with considerable confidence.  We can
54: improve our confidence by repeating the experiment, but for
55: simplicity, let us suppose that $\epsilon$ is so miniscule that we are
56: satisfied with doing a single experiment.
57: 
58: In drawing the above conclusions, we acted as if our laboratory either
59: was the only laboratory in the universe, or was selected at random
60: from among all the laboratories doing the same experiment in the
61: universe.  This is the assumption of typicality.  Note that we have no
62: direct evidence for this assumption.  We do not know whether there are
63: other laboratories performing the same experiment on some far-away
64: planets; and if there are, then our laboratory was presumably not
65: actually selected by anyone from among them.  Nevertheless, the
66: overall success of the scientific method so far suggests that this
67: assumption is appropriate.
68: 
69: To see this, consider a prescription favored by Hartle and
70: Srednicki~\cite{HarSre07}, who decline to assume typicality.  They argue
71: that it does not matter whether a given outcome is likely to occur in
72: a randomly chosen laboratory; what matters is whether one is likely to
73: be able to find {\em some\/} laboratory, somewhere in all of
74: spacetime, no matter how atypical, in which that outcome occurs.  This
75: probability is given not by $\epsilon$, but by $1-(1-\epsilon)^L$,
76: where $L$ is the number of laboratories in the universe.
77: 
78: The effect of using this probability-of-global-existence is most
79: dramatic in the case where $L\gg \epsilon^{-1}\gg 1$.  Then we cannot
80: rule out either theory, no matter what we observe.  We can still rule
81: out one of the two theories by repeating the experiment sufficiently
82: often.  But to know at which point we can reject one of the theories,
83: we would need to know how many other laboratories there are.  Since we
84: do not know $L$, the Hartle-Srednicki prescription would put an end to
85: experimental science.  It would render all experiments pointless,
86: because we could not reject any theory until we know how many other
87: laboratories there are.  Given the success of the scientific method
88: thus far, we may conclude the Hartle-Srednicki prescription is
89: inappropriate.\footnote{We cannot conclude that our laboratory is the
90:   only one, since we could simply build a second one.  Note that in
91:   the Hartle-Srednicki prescription, this would be inadvisable, since
92:   it would render the experiments performed at either laboratory less
93:   conclusive.}
94: 
95: Here we have argued for the assumption of typicality on empirical
96: grounds: it has served us well as a heuristic tool.  If it was wrong,
97: we should not have been successful in devising and rejecting
98: scientific theories on the basis of this assumption.  But {\em why}\/
99: does it work so well?  This, too, can be understood; elegant
100: discussions have recently been given by Page~\cite{Pag07}, and by
101: Garriga and Vilenkin~\cite{GarVil07}, who also offer a careful
102: definition of the class of observers among which we may consider
103: ourselves to be typical.
104: 
105: 
106: \subsection{The measure problem: a phenomenological approach}
107: 
108: In the multiverse, we can use typicality to make statistical
109: predictions for the results of observations. For instance, to predict
110: the cosmological constant, we would first determine the theoretically
111: allowed values, and then count the number of observations of each
112: value.  The probability to observe a given value of the cosmological
113: constant is proportional to the number of observations, in the
114: multiverse, of that value.  The problem is that under rather generic
115: conditions, the universe will have infinite spacetime volume, even if
116: it is spatially finite (i.e., contains a compact Cauchy surface).
117: Then the number of observations can diverge.
118: 
119: The landscape of string theory contains perhaps $10^{500}$ metastable
120: vacua, allowing it to solve the cosmological constant
121: problem~\cite{BP}; see Refs.~\cite{Pol06,TASI07} for a review.
122: However, divergences would arise even if there was only one false
123: vacuum.  For example, suppose that there was a first-order phase
124: transition in our past, by which a long-lived metastable vacuum
125: decayed.  The symmetries of the instanton mediating this
126: decay~\cite{CDL} dictate that the resulting true vacuum region is an
127: infinite open FRW universe.  It will contain either no observers, or
128: an infinite number of them.  Moreover, the parent vacuum will keep
129: expanding faster than it decays, so that an infinite number of true
130: vacuum bubbles (or ``pocket universes'') are created over
131: time~\cite{GutWei83}.
132: 
133: The measure problem in cosmology is the question of how to regulate
134: these infinities, in order to get a finite count of the number of
135: observations of each type.\footnote{In general, the question of what
136:   constitutes one observation is a difficult problem.  For instance,
137:   it is not obvious precisely how many observations of the CMB
138:   temperature should be assigned to our local efforts on Earth.  (See
139:   Ref.~\cite{Bou06} for a recent proposal using entropy production.)
140:   However, these considerations are orthogonal to the issue at hand,
141:   which is the regularization of the infinite spacetime four-volume
142:   arising in eternal inflation.  Here, we will assume that the local
143:   counting of observations is unambiguous.}  The choice of measure is
144: no minor technicality, but an integral part of a complete theory of
145: cosmology.  Two different measures often assign exponentially
146: different relative probabilities to two types of
147: observations.\footnote{A simple example arises even if there is only
148:   one false vacuum.  Each true vacuum bubble collides with an infinite
149:   number of other such bubbles, so one may ask whether we are likely
150:   to live in a collision region.  Leaving aside fatal effects of
151:   collisions, this probability is nevertheless exponentially small in
152:   the proper time measure considered here, and also in the causal
153:   diamond measure~\cite{Bou06}.  But if one averages over worldlines
154:   emanating from the nucleation point, all but a set of measure zero
155:   of them will immediately enter a collision region~\cite{GarGut06}.}
156: 
157: % For the counting of observations to have predictive power, we must
158: % assume that in a sense our observations are chosen at random among all
159: % the possible observations which are made in the multiverse. (One could
160: % also focus on a subset of observations by asking a more restricted
161: % question like ``What is the expected cosmological constant given all
162: % other known physics?''  But again the statistical approach would
163: % assume that our observations are chosen at random among all of the
164: % observations consistent with the prior assumptions.)  This {\em
165: %   typicality} assumption has been criticized in the multiverse context
166: % but, as we discuss in Section 2, it is used every day when we use
167: % error analysis to compute the statistical significance of experimental
168: % results. We further point out that a proposed modification \cite{HS}
169: % of the usual typicality assumption wreaks havoc on the scientific
170: % method, by requiring experimentalists to determine the number of labs
171: % in the multiverse in order to compute the statistical significance of
172: % their results.
173: 
174: Ultimately, a unique measure should arise from first principles in a
175: fundamental theory~\cite{FreSus04,FreSek06,Sus07,MalSheSus}.  
176: In the meantime, however, we may regard the measure problem as a {\em
177: phenomenological\/} challenge.  At least in the semiclassical
178: regime, we can hope to identify the correct measure by the traditional
179: scientific method: We try a simple, minimal theory, and work out its
180: implications.  If they conflict with observation, we either refine
181: (i.e., complicate) the model, or we abandon it altogether for a
182: different approach.
183: 
184: What one may regard as a simple measure is, to some extent, in the eye
185: of the beholder.  The same can be said for simple theories; yet, for
186: the most part, we know one when we see one.  Only a handful of
187: measures have been proposed (see, e.g, Refs.~\cite{Vil06,Lin06,Van06}
188: for overviews and further references),
189: %GarSchVil~\cite{LinLin94,GarLin94,GarVil01,GarSch05,Bou06,everybody},
190: and many of them can be seen to conflict with observation, often
191: violently.  This is good news, because it makes it feasible to proceed
192: by elimination.  Let us investigate simple proposals, let us ask
193: whether they are well-defined, and let us determine whether they
194: conflict with observation.
195: 
196: For example, consider the proposal of Ref.~\cite{GarSch05}.  In its
197: original form, it predicted with probability 1 that we should find
198: ourselves as isolated observers (``Boltzmann brains'') resulting from
199: a highly suppressed thermal fluctuation in a late, empty
200: universe~\cite{Pag06,BouFre06b}.  This led to a
201: refinement~\cite{Vil06b}, which complicates the measure and seems {\em
202:   ad hoc}~\cite{Pag06b}.  Depending on the details of the string
203: landscape, the proposal may render most vacua dynamically inaccessible
204: (the ``staggering problem'' of Refs.~\cite{SchVil06,OluSch07}).  This
205: would also amount to a conflict with observation, namely the
206: prediction that we should observe a much larger cosmological constant
207: with probability very close to 1.  Perhaps most importantly, at
208: present the proposal is well-defined only in the thin-wall limit of
209: bubble formation, and if bubble collisions are
210: neglected~\cite{GarGut06}.\footnote{The proposal cuts off the infinite
211:   number of observers in different vacuum bubbles by restricting to a
212:   ``unit comoving volume'', defined by appealing to the universality
213:   of the open universe metric inside every bubble at early
214:   times~\cite{GarSch05}.  But universality holds only if the thickness
215:   of the wall, and its collisions with other bubbles, are both
216:   neglected.  These two assumptions cannot both be satisfied to a good
217:   approximation.}
218: 
219: Another recent proposal~\cite{Bou06}, the ``holographic'' or ``causal
220: diamond'' measure, has so far fared well.  It is
221: well-defined in the semiclassical limit, and it does not have a
222: staggering problem~\cite{BouYan07}.  Its prediction of the
223: cosmological constant agrees significantly better with the data than
224: that of any other proposal~\cite{BouHar07}, and it continues to agree
225: well even as other parameters are allowed to vary~\cite{CliFre07}.  It
226: will be important to test this proposal further, for example, by
227: allowing even more parameters to vary.  But it is encouraging that we
228: have at least one well-defined measure that has not been ruled out.
229: 
230: In this paper, we consider a much older proposal, the {\em proper time
231:   measure\/}~\cite{Lin86a,LinLin94,GarLin94,GarLin94a,GarLin95}.  At
232: present, this measure is not completely well-defined, and we will
233: comment on some issues that will have to be overcome to make it
234: well-defined.  But our main focus will be on its well-known conflict
235: with observation, the ``youngness paradox''.  In particular, we will
236: investigate whether simple modifications of the measure can resolve
237: this problem.
238: 
239: 
240: \subsection{The proper time measure and the youngness paradox}
241: 
242: To apply the proper time measure, one begins by selecting an (almost
243: arbitrary) finite portion of a spacelike slice in the semiclassical
244: geometry.  The congruence of geodesics orthogonal to this initial
245: surface defines Gaussian normal coordinates, and thus a time slicing,
246: at least until caustics are encountered. The number of observations
247: between the initial slice and the time $t$ is finite.  Globally, the
248: multiverse reaches a self-reproducing state at late times: its volume
249: expands exponentially, but the ratio of different types of
250: observations remains constant and finite.  Therefore, relative
251: probabilities defined by this measure are independent of the initial
252: conditions.
253: 
254: Earlier work~\cite{LinLin96,Gut00a,Gut00b,Gut04,Teg05,Lin07,Gut07} has 
255: already shown that the proper time measure has a youngness problem: it
256: predicts with essentially 100\% probability that we should be living
257: at an earlier time.  The reason for this problem can roughly be
258: described as follows.
259: 
260: The asymptotic rate of expansion of the multiverse is dominated by the
261: vacuum with the largest Hubble constant $H_{\rm big}$, which defines a
262: microphysical time-scale $H_{\rm big}^{-1}$.  (In the string
263: landscape, this would be of order the Planck time.)  For simplicity,
264: let us consider only regions occupied by our own vacuum.  We may ask
265: about the distribution of the age of such bubbles, i.e., how long
266: before the cutoff $t$ they were formed.  In particular, we may ask how
267: many bubbles are at least $13.7$ Gyr old, and thus contain
268: observations like ours; and we may compare this to the number of
269: bubbles that are, say, $13$ Gyr old.  The size of the bubble interior
270: is not much affected by these different time choices, but the number
271: of bubbles will be vastly different.  For every bubble that is at
272: least $13.7$ Gyr old at the time $t$, there will be of order $\exp
273: \left(3\times 0.7 \,{\rm Gyr}/ H_{\rm big} ^{-1}\right)$ bubbles that
274: are $13$ Gyr old, because of the overall exponential growth of the
275: volume of the multiverse in the extra 700 million years before it has
276: its last chance to nucleate the younger bubbles.  Perhaps the younger
277: bubbles contain fewer observers per bubble, but surely not so few as
278: to compensate for a factor $\exp(10^{60})$.  This mismatch persists as
279: $t\to\infty$. Thus, typical observers are younger than we are, and the
280: probability for an observer to live as late as we do is
281: $\exp(-10^{60})$.  This rules out the proper time measure at an
282: extremely high level of confidence.
283: 
284: Of course, our choice of $13$ Gyr observers as a comparison group is
285: arbitrary.  Because $H_{\rm big}^{-1}$ is a microphysical scale, even
286: observers just one minute younger (relative to their big bang) are
287: superexponentially more probable than we are.  Ultimately, one should
288: consider observers of any cosmological age.  Because of the
289: exponential pressure to be young, it pays to arise from a rare quantum
290: fluctuation in the early universe.  The most likely observers are such
291: ``Boltzmann babies'', and the most likely observations are the
292: phenomena of the hot, dense, early universe they see.
293: 
294: 
295: \subsection{Summary and outline}
296: 
297: Our goal in this paper is two-fold.  First, we will make the youngness
298: paradox more precise.  Traditional treatments have neglected the
299: expansion of new bubbles.  We supply a justification for this ``square
300: bubble'' approximation, by extending Gaussian normal coordinates
301: across an expanding bubble wall, and showing that our exact treatment
302: reproduces the usual youngness problem.  We will distinguish
303: carefully between probability distributions for the {\em time\/} when
304: observers live (which is not directly observable), and probability
305: distributions for actual observables, like the temperature of the
306: background radiation measured by observers~\cite{Teg05}.  We find that
307: the youngness paradox manifests itself by predicting that we should
308: observe a higher temperature than 2.7 K, with probability
309: exponentially close to 1.
310: 
311: Our second goal is to consider possible modifications of the proper
312: time measure.  We will argue that it is difficult to resolve the
313: youngness paradox, other than by abandoning the measure altogether.
314: In particular, Linde has proposed a modification in the context of a
315: particular toy model~\cite{Lin07}. Since no general prescription was
316: given, it is not clear how to extend this modification to other
317: settings, and in particular to the probability distribution for the
318: observed background temperature.  We consider a number of possible
319: choices, some of which reduce to the prescription of Ref.~\cite{Lin07}
320: for the particular probabilities computed therein.  However, we are
321: unable to find any modification that escapes all of the conflicts with
322: observation that arise from the youngness problem.  In particular, 2.7
323: K remains an extremely atypical value of the background temperature
324: under all choices we consider.
325: 
326: The structure of the paper is as follows.  In Sec.~\ref{sec-proper},
327: we explain the proper time measure in more detail. In
328: Sec.~\ref{sec-geod}, we compute the paths of geodesics entering
329: bubbles, in order to determine the shape of the proper time cutoff
330: within bubbles. In Sec.~\ref{sec-prob}, we compute the probability
331: distribution for the spacetime location of observers, finding a
332: youngness paradox and conflict with observation.  In
333: Sec.~\ref{sec-fixes}, we try a few modifications of the measure, but
334: find no simple modification consistent with observation.
335: 
336: \section{The Proper Time Measure}
337: \label{sec-proper}
338: 
339: The proper time measure (sometimes referred to as the ``standard
340: volume weighted measure'') is one of the simplest and most
341: straightforward ways of regulating the infinities of the multiverse.
342: Choose a small three-dimensional patch of space, $\Sigma_0$,
343: orthogonal to at least one eternally inflating geodesic.  Then,
344: construct Gaussian normal coordinates~\cite{Wald} in its future.  That
345: is, a given event has the time coordinate $t$ if it occurs at proper
346: time $t$ along a geodesic orthogonal to $\Sigma_0$.  Such events form
347: a three-dimensional hypersurface $\Sigma_t$.  The regularization
348: scheme is to count only observations between proper time hypersurfaces
349: $\Sigma_0$ and $\Sigma_t$.  Relative probabilities are defined by
350: ratios, in the limit $t\rightarrow\infty$.
351: 
352: \begin{figure}[h!]
353: \begin{center}
354: \includegraphics[viewport=600 340 0 0,width=8 cm,clip,angle=180]{multi}
355: \caption{The relative probability of making different observations,
356:   for example two different CMB temperatures (red disks or blue
357:   boxes), is determined by simple counting in the finite region
358:   between $\Sigma_0$ and $\Sigma_t$.  The ratio tends to a finite
359:   limit as $t\rightarrow\infty$.  The youngness problem is the fact
360:   that anomalous early fluctuations producing either observation
361:   (Boltzmann babies) turn out to dominate the count.  To show this
362:   correctly in the figure, one would need to draw an exponentially
363:   large number of ``young bubbles'', like the one on the right, in
364:   which only the Boltzmann babies contribute.}
365: \label{fig-multi}
366: \end{center}
367: \end{figure}
368: 
369: It is well-known that Gaussian normal coordinates are only locally
370: defined.  They break down at {\em caustics}, or focal points, where
371: infinitesimally neighboring geodesics in the congruence intersect.
372: Beyond such points, the above definition of the time coordinate $t$ is
373: ambiguous.  We sidestep the issue here by considering only expanding
374: spacetime regions and ignoring clustering and inhomogeneities (and
375: thus, strictly speaking, all known observers), so that focusing does
376: not occur.
377: 
378: Let $O_1$ and $O_2$ be two mutually exclusive observations.  For
379: example, $O_1$ may subsume any observation made in vacuum $A$, while
380: $O_2$ corresponds to vacuum $B$.  Or $O_1$ ($O_2$) may be capture
381: information about the observer's spatial or temporal location within a
382: given vacuum, for example the fact that universe is matter (vacuum)
383: dominated.
384: 
385: Let $N_i(t)$ be the number of observations of type $O_i$ made in the
386: four volume between $\Sigma_0$ and $\Sigma_t$.  Observations
387: take a finite time, so for definiteness let us demand that an
388: observation must be complete for it to be counted.  The relative
389: probability for the two observations is defined to be
390: \begin{equation}
391: \frac{p(O_1)}{p(O_2)}=\lim_{t\to\infty}\frac{N_1(t)}{N_2(t)}~.
392: \label{eq-prob}
393: \end{equation}
394: 
395: Similarly, we can consider a continuous set of possible observations
396: $O_T$, such as the observation of a CMB temperature $T$.  In this
397: case, we are interested in the probability density $dp/dT$, which is
398: given by
399: \begin{equation}
400:   \frac{\left.\frac{dp}{dT}\right|_{T_1}
401:   }{\left.\frac{dp}{dT}\right|_{T_2}}=
402:   \lim_{t\to\infty}
403:   \frac{\left.\frac{dN}{dT}\right|_{T_1}(t)
404:   }{\left.\frac{dN}{dT}\right|_{T_2}(t)}
405: \label{eq-density}
406: \end{equation}
407: Here, $\left.\frac{dp}{dT}\right|_{T_1}dT$ is the
408: probability of observing $T$ in the interval
409: $(T_1,T_1+dT)$.
410: $\left.\frac{dN}{dT}\right|_{T_1}dT$ is the number of
411: instances of such observations in the four-volume between $\Sigma_0$
412: and $\Sigma_t$.\footnote{In the continuous case, one must take care
413:   with the order of limits.  First we pick a finite $dT$ and define
414:   ratios as usual, by taking $t\to\infty$.  Then we repeat the
415:   procedure while taking $dT\to 0$.}
416: 
417: At late times, $N_i(t)\propto \exp(3H_{\rm big}t)$.  The overall
418: scaling rate $H_{\rm big}$ is set by the most rapidly expanding
419: vacua~\cite{Linde}.  [In the string landscape, one expects that
420: $H_{\rm big}\sim {O}(1)$ in Planck units.]  This exponential
421: growth guarantees that
422: \begin{equation}
423:   \lim_{t\to\infty}\frac{N_1(t)}{N_2(t)}=
424:   \lim_{t\to\infty}\frac{N_1'(t)}{N_2'(t)}~,
425: \end{equation}
426: where $N_i'=dN_i/dt$ is the rate at which observations of type $O_i$
427: are being made, integrated over space but not over time.  Thus, it
428: does not matter whether probabilities are computed from the total
429: number of observations until the time $t$, or the rate of observations
430: at the time $t$, or the number of observations made in some recent
431: (fixed width) time interval $(t-\Delta t,t)$.  For definiteness,
432: however, we will stick to the first of these definitions.
433: 
434: 
435: 
436: \section{Geodesics crossing bubbles}
437: \label{sec-geod}
438: \subsection{Open FRW time vs. geodesic time}
439: 
440: The measure discussed above was first applied to slow-roll models of
441: eternal inflation, without first-order phase
442: transitions~\cite{LinLin94,GarLin94}.  In this case, one keeps track
443: of fluctuations of scalar fields on the Hubble scale, effectively
444: assuming that they decohere every Hubble time (see
445: Ref.~\cite{BouFre06a} for a discussion of the validity of this
446: approach).  There is no obstruction to applying the same measure to
447: models with bubble formation, but there is an annoying complication
448: (Fig.~\ref{fig-slic}).
449: 
450: \begin{figure}[t!]
451: \begin{center}
452: \includegraphics[scale=1]{slic}
453: \includegraphics[scale=1]{geod}
454: \caption{The top figure shows slices of constant FRW time, $\tau$
455:   (red, light) and slices of constant geodesic time $t$ (blue, dark)
456:   in the vicinity of a bubble wall (green, thick) with initial size
457:   $r_0=0.1\, H_{\rm out}^{-1}$.  Note that the constant $t$ slices are
458:   not defined for geodesics passing through the nucleation region of
459:   the bubble.  The lower figure shows that geodesics of the congruence
460:   (blue, dark) eventually asymptote to comoving FRW worldlines (red,
461:   light).}
462: \label{fig-slic}
463: \end{center}
464: \end{figure}
465: 
466: Consider a region of the universe at late times, occupied by a
467: ``host'' de~Sitter vacuum with cosmological constant $3H_{\rm out}^2$.
468: Let us suppose that $H_{\rm out}$ is very large, but far enough below
469: the Planck scale to lend validity to our semiclassical treatment.
470: Moreover, we suppose that a bubble of our own vacuum can form by a
471: Coleman-DeLuccia (CDL) tunneling process inside the host vacuum.
472: 
473: Let us suppose, moreover, that the host vacuum has existed for many
474: Hubble times $H_{\rm out}^{-1}$.  Then, on the scale about to be
475: occupied by a newly formed bubble, the geodesics emanating from the
476: initial surface $\Sigma_0$ can be treated as comoving in the flat
477: de~Sitter metric
478: \begin{equation}
479: ds^2=-dt^2+ H_{\rm out}^{-2} e^{2H_{\rm out}t} d {\mathbf x} ^2
480: \label{eq-flat}
481: \end{equation}
482: This follows, in a sense, from the de~Sitter no-hair theorem; we will
483: also find that it is consistent with our careful analysis in
484: Sec.~\ref{sec-exact}. 
485: 
486: Now suppose that a bubble of our vacuum forms at the time $t_{\rm nuc}
487: $.  It will appear at rest, with a proper radius $r_0\ll H_{\rm
488:   out}^{-1}$ determined by the CDL instanton.  Then it will expand at
489: constant acceleration $r_0^{-1}$, its world-volume asymptoting to a
490: light-cone.  Some of the above geodesics will eventually run into the
491: bubble wall and enter our universe.  Their behavior will
492: determine the weight of any observations carried out inside the
493: bubble, in the proper time measure.
494: 
495: We will not consider the
496:   small subset of geodesics that go through the nucleation region,
497:   $r\lesssim r_0$, where the classical geometry is not clearly
498:   defined.  It is unclear how to treat these geodesics.  This
499:   constitutes a challenge for the sharp formulation of
500:   congruence-based measures.  The best we can say is that our results
501:   show that values of $r\sim O(r_0)$ contribute negligibly to the
502:   measure as they are approached from above in a controlled regime.
503:   This could be viewed as evidence that the contribution of the
504:   uncontrolled regime can also be neglected.
505: 
506: The metric inside the bubble is given by an open FRW geometry;
507: ignoring fluctuations, the metric is
508: \begin{equation}
509: ds^2=-d\tau^2+a(\tau)^2 (d\xi^2+\sinh^2\xi\, d\Omega_2^2)~.
510: \label{eq-FRW}
511: \end{equation}
512: where the scale factor $a(\tau)$ comprises, for anthropically
513: relevant bubbles, a period of
514: inflation followed by radiation, matter, and vacuum domination with
515: very small cosmological constant.  Note that
516: $a(\tau)\approx \tau$ for sufficiently small $\tau$.
517: 
518: The maximally symmetric and negatively curved spatial slices
519: defined by $\tau = const$ are physically preferred inside the bubble, since they
520: correspond to hypersurfaces of (approximately) constant density.  Only
521: at very late times, well into the vacuum-dominated era, do we lose
522: this preferred slicing, as the universe again becomes locally empty
523: de~Sitter.
524: 
525: The key point is that the preferred surfaces of constant FRW time $\tau$
526:  are {\em  not\/} the surfaces  $\Sigma_t$ of constant geodesic time $t$.
527:     This is a complication, since
528: $\tau$ is what we usually call the age of the universe, the time since
529: the big bang---really, the metric distance from the bubble nucleation
530: event.  To the extent that any time variable is directly correlated
531: with the outcome of an observation (such as CMB temperature or the
532: amount of clustering), that variable will be the FRW time $\tau$, 
533: and not the global time $t$.
534: 
535: \subsection{Square bubble approximation}
536: \label{sec-square}
537: 
538: The {\em square bubble approximation\/}, which is implicit in
539: Ref.~\cite{Lin06}, aims to circumvent this complication.  It
540: amounts to a deformation of the metric that allows us to calculate as
541: if constant $\tau$ slices, inside the bubble, coincide with constant
542: $t$  slices.  For this we must arrange that the FRW
543: time $\tau$ and the geodesic time $t$ differ only through a
544: constant shift,
545: \begin{equation}
546: \tau=t-t_{\rm nuc}~.
547: \end{equation}
548: This is possible only if the movement of the bubble wall is neglected.
549: 
550: Given this {\em ad-hoc\/} modification, the continuation of the
551: geodesic congruence into the bubble cannot be directly computed.  We
552: will simply assume that the internal geometry of the new vacuum is a
553: spatially finite piece of a {\em flat\/} FRW universe
554: \begin{equation}
555:   ds^2=-d\tau^2 + \tilde a(\tau)^2 d {\mathbf y} ^2~.
556: \label{eq-insideflat}
557: \end{equation}
558: To match at $t=t_{\rm nuc} $ ($\tau=0$), we let $ {\mathbf y}$ range
559: over a finite physical volume $\tilde a(0)^3 V_y$.  We take the
560: comoving volume to be independent of $\tau$, as if the bubble wall
561: remained at fixed ${\mathbf y} $.
562: 
563: Note that both the scale factor and the initial size of the bubble
564: initially differ significantly from their true values, and the
565: matching to the outside fails at late times.  However, in inflating
566: vacua the exponential internal growth is more important than the
567: expansion of the bubble forming their boundary.  Moreover, inflation
568: locally washes out the difference between a flat and an open universe.
569: After a short time (say, a few e-foldings of inflation) we can take
570: $\tilde a(\tau)\propto a(\tau)$.
571: 
572: Nevertheless, the square bubble approximation blatantly contradicts
573: important known features, such as the fact that the constant-density
574: slices inside are actually open and infinite.  Indeed it is not even
575: consistent geometrically, making it impossible to match the inside of
576: the bubble to the outside.  But one may hope that it gives a
577: reasonable approximation {\em for the purpose of computing
578:   probabilities}.  This will be the case if the approximation does not
579: change the true count of observations of various types. We will find
580: that the square bubble approximation is a good one for many questions.
581: 
582: 
583: \subsection{Exact relation}
584: \label{sec-exact}
585: The actual relation between the FRW coordinates $(\tau,\xi)$ and the
586: geodesic proper time $t$ is more complicated. We set
587: \begin{equation}
588: H_{\rm out} = 1 \ \ {\rm in\ this\ subsection,}
589: \end{equation}
590:  so that the equations are not quite so ugly.
591: In the outside flat deSitter slicing, 
592: \begin{equation}
593: ds^2 = - dt^2 + e^{2 t } \left(dr^2 + r^2 d \Omega_2^2 \right) ~,
594: \end{equation}
595:  the domain wall 
596: follows the trajectory
597: \begin{eqnarray}
598: r_w \exp (t_w - t_{\rm nuc})&=&r_0\cosh\eta~, \\
599: \exp (t_w - t_{\rm nuc})&=&
600: r_0\sinh\eta+\sqrt{1-r_0^2}~. \label{eq-tw}
601: \end{eqnarray}  
602: Here $r_0$ is the size of the bubble at nucleation; it is also the
603: radius of curvature and the inverse proper acceleration of the domain
604: wall.  $r_0 \eta$ is the proper time along the domain wall.
605: 
606: We need to compute the motion of geodesics as they cross the domain
607: wall and live happily ever after in the interior. The natural
608: coordinates inside the bubble are the open FRW coordinates $(\tau,
609: \xi)$ of Eq.~(\ref{eq-FRW}) because they respect the symmetry of the
610: bubble nucleation. However, these coordinates do not cover the region
611: containing the domain wall, so it is convenient to use a different
612: coordinate system near the domain wall.  Assuming that the Hubble constant 
613: in the interior of the bubble is much smaller than the Hubble constant in 
614: the exterior, we can find a scale $\tau^*$ such that 
615: $H_{\rm in}^{-1}\gg\tau^*\gg H_{\rm out}^{-1}$.  
616: The region from the domain wall to the 
617: $\tau^*$ surface is much smaller than the characteristic scales of the geometry inside the bubble.  As a result, we can approximate it as a piece of Minkowski space. We will use coordinates in which the metric is
618: \begin{equation}
619: ds^2 = - dT^2 + dR^2 + R^2 d\Omega_2^2~.
620: \end{equation}
621:  
622: Because the domain wall is a constant curvature surface with curvature
623: radius $r_0$, its trajectory in the Minkowski coordinates is
624: \begin{eqnarray}
625: R_w&=&r_0\cosh\eta~, \\
626: T_w&=&r_0\sinh\eta~,
627: \end{eqnarray} 
628: where again $r_0 \eta$ is the proper time along the domain wall.
629: Computing the 4-velocity, we find that $\eta$ is the rapidity of the
630: domain wall.
631:   
632: The trajectory of the geodesic after crossing the domain wall is
633: \begin{eqnarray}
634: T&=&r_0\sinh\eta+(t-t_w)\cosh\alpha \label{eq-T}~, \\
635: R&=&r_0\cosh\eta+(t-t_w)\sinh\alpha \label{eq-R}~,
636: \end{eqnarray}
637: where $\alpha$ is the rapidity of the geodesic.  We will determine
638: $\alpha$ by demanding that the angle between the domain wall and the
639: geodesic is continuous across the domain wall\footnote{At the domain
640:   wall, the first derivative of the metric is discontinuous, resulting
641:   in a delta function in $R_{\mu\nu}$ in the thin wall
642:   approximation. However because the connection only depends on first
643:   derivatives, we can find a local coordinate system where the
644:   connection $\Gamma^\gamma_{\mu\nu}$ is finite.  This proves that the
645:   angle (inner product) between the geodesic and the domain wall is
646:   continuous across the wall.}. If $u$ and $v$ are the 4-velocities of
647: the geodesic and the domain wall, we demand
648: \begin{equation}
649: u\cdot v|_{\rm out} =u\cdot v|_{\rm in} ~.
650: \end{equation}
651: Since $\alpha$ is the rapidity of the geodesic and $\eta$ is the
652: rapidity of the domain wall,
653: \begin{equation}
654: u\cdot v|_{\rm in} =\cosh(\eta - \alpha)~.
655: \end{equation}
656: Geodesics outside the domain wall have a simple 4-velocity $u_{\rm
657:   out} = (1, 0, 0, 0)$, and since we have identified $r_0 \eta$ as the
658: proper time along the domain wall, the 4-velocity of the domain wall
659: is $v = ({1 \over r_0} ~ d t_w /d \eta, ...)$. Using the equation
660: (\ref{eq-tw}) for the trajectory of the domain wall,
661: \begin{equation}
662: u\cdot v|_{\rm out}
663: ={1 \over r_0} \frac{d t_w }{d \eta}
664: =\frac{\cosh\eta}{r_0\sinh\eta+\sqrt{1-r_0^2}}~.
665: \end{equation}
666: Thus the equation determining $\alpha$ is
667: \begin{equation}
668: \frac{\cosh\eta}{r_0\sinh\eta+\sqrt{1-r_0^2}}
669: =\cosh(\eta-\alpha)~.
670: \label{eq-angle}
671: \end{equation}
672: 
673: It is convenient to combine Eq.~(\ref{eq-angle}) and (\ref{eq-tw})
674: to get
675: \begin{equation}
676: \cosh \eta \exp \left[-(t_w - t_{\rm nuc})\right] = \cosh(\eta - \alpha)~.
677: \end{equation}
678: Simplifying we find
679: \begin{equation}
680: (t_w - t_{\rm nuc})=\alpha-\ln(1+\varepsilon)~,
681: \label{eq-raptw}
682: \end{equation}
683: where
684: \begin{equation}
685: \varepsilon=\frac{e^{2 \alpha} - 1}{e^{2 \eta} + 1}~.
686: \label{eq-eps}
687: \end{equation}
688: This is a convenient rewriting because one can show that $\varepsilon
689: \ll 1$ for all geodesics as long as the critical bubble size is small
690: in Hubble units, $r_0 \ll 1$.
691: 
692: We want to rewrite the geodesics in terms of the open FRW coordinates
693: which will be adapted to the cosmological evolution inside the
694: bubble. For $\tau\ll H_{\rm in}^{-1}$, where the geometry is approximately
695: Minkowski space, the relationship is
696: \begin{eqnarray}
697: \tau &=& \sqrt{T^2-R^2} \label{eq-tau} \\
698: \xi  &=& \tanh^{-1}\frac{R}{T} \label{eq-xi}~.
699: \end{eqnarray}
700: Using the trajectory in $(R,T)$ given by Eq.~(\ref{eq-T}), (\ref{eq-R}), 
701: we find the trajectory in FRW coordinates
702: \begin{eqnarray}
703: \xi&=&\alpha+{O}(\frac{1}{\tau})~,
704: \label{eq-xia}
705: \\
706: \tau &=&\sqrt{(t - t_w)^2+2r_0(t-t_w)\sinh(\eta-\xi)-r_0^2}~.
707: \label{eq-taua}
708: \end{eqnarray}
709: Our goal is to manipulate all of the above equations in order to find
710: a single equation for the geodesic time since nucleation, $ t - t_{\rm
711:   nuc}$, as a function of the natural coordinates $\tau, \xi$ inside
712: the bubble.
713: 
714: We will be interested in events which occur a reasonable distance away
715: from the domain wall, so that $\tau, t-t_{\rm nuc}, t-t_w \gg 1 $. So
716: we can drop the subleading terms in (\ref{eq-xia}) and just set the
717: rapidity of the geodesic equal to the comoving coordinate, $\alpha =
718: \xi$. Physically, the point is that the final comoving position of the
719: geodesic is determined only by its velocity and not by its initial
720: location. The nontrivial statement in (\ref{eq-xia}) is that the
721: geodesics become comoving in a time set by the Hubble scale outside
722: the bubble, $H_{\rm out}^{-1}$.
723:  
724: In Eq.~(\ref{eq-taua}) we can now set $\alpha = \xi$ and expand for
725: large $t-t_w $ to get
726: \begin{equation}
727: \tau =t-t_w + r_0 \sinh(\eta - \xi)~.
728: \end{equation}
729: Going back to (\ref{eq-angle}) and solving for $\sinh(\eta - \xi)$ we find
730: \begin{equation}
731: \sinh(\eta - \xi) = \frac{\sqrt{1-r_0^2}\sinh\eta-r_0}
732: { r_0 \sinh\eta+\sqrt{1-r_0^2}}~.
733: \end{equation}
734: Using the relation (\ref{eq-tw}) between $\eta$ and $t_w$ this can be
735: rewritten as
736: \begin{equation}
737:   \sinh(\eta - \xi) = {1 \over  r_0} \left[\sqrt{1 -  r_0^2}
738:     - \exp \left[-(t_w-t_{\rm nuc})\right] \right]~.
739: \end{equation} 
740: So we have an equation relating the geodesic time since nucleation to
741: the FRW time $\tau$ and the time $t_w$ the geodesic crosses the domain
742: wall:
743: \begin{equation}
744:   \tau = t - t_w +   \left[\sqrt{1 -  r_0^2} - 
745:     \exp[-(t_w-t_{\rm nuc})] \right]~.
746: \end{equation}
747: 
748: Now we can use the relation (\ref{eq-raptw}) between the rapidity
749: $\alpha$ and $t_w$, together with $\alpha= \xi$, to get
750: \begin{equation}
751: \tau = t - t_{\rm nuc}-\bigg[\xi + e^{- \xi} -
752: \sqrt{1-r_0^2}+ \varepsilon e^{-\xi} -\ln(1+\varepsilon)\bigg]~,
753: \label{eq-exact}
754: \end{equation}
755: where $\varepsilon$ is given by (\ref{eq-eps}).  Expanding in
756: $\varepsilon$ and restoring the factors of $H_{\rm out}$, we get the
757: final formula relating the geodesic time to the natural coordinates
758: inside the bubble:
759: \begin{equation}
760:   t - t_{\rm nuc} = \tau + H_{\rm out}^{-1}
761:   \left[ \xi + e^{- \xi} - 1 + ... \right]~.
762: \label{eq-app}
763: \end{equation}
764: As expected, the difference between the geodesic proper time and the
765: open FRW time depends non-trivially on the radial FRW coordinate
766: $\xi$.
767: 
768: \section{The spacetime location of a typical observer}
769: \label{sec-prob}
770: 
771: The proper time measure makes nontrivial and interesting predictions
772: for vacuum selection, which do not appear to contradict anything we
773: know~\cite{CliShe07}.  However, as soon as we ask about the
774: probabilities of different observations in the same vacuum, the
775: measure wildly conflicts with observation.  It has two properties that
776: result in a squeeze.  On the one hand, for an observation to be
777: counted, it must occur before the cutoff $t$.  On the other hand, the
778: multiverse as a whole is expanding exponentially on a microscopic
779: characteristic time scale.  This makes it favorable to wait as long as
780: possible until creating a low-energy, slowly expanding region like the
781: one in which we are making our observations, and it strongly favors
782: observations that happen soon after the fastest expanding vacuum has
783: decayed.  This is the general idea of the youngness paradox
784: ~\cite{LinLin96,Gut00a,Gut00b,Gut04,Teg05,Lin07,Gut07}.
785: 
786: We will present one explicit calculation to show the fact that, within 
787: bubbles identical to ours, the probability to live at $13.7$ Gyr is 
788: vanishingly small compared to the probability to live at $13$ Gyr.  
789: There is nothing new in this calculation, but it will be easier to see
790: how the exact geometry we found goes into the paradox in Sec.
791: \ref{sec-bubble}, and how to analyze possible modifications in Sec.
792: \ref{sec-fixes}.  
793: 
794: Another manifestation of the youngness paradox is that if a number of
795: tunneling events are necessary to get from the fastest inflating
796: vacuum to our host vacuum, these successive tunneling events will tend
797: to be separated by only the Planckian time interval $H_{\rm
798:   big}^{-1}$. Since the tunneling events are not well-separated, this
799: renders it difficult to compute semiclassically. However, since such a
800: quick succession of tunneling events does not obviously contradict
801: observation, we sidestep this difficulty here by assuming that our
802: vacuum is produced directly from the fastest inflating vacuum. Hence
803: we set 
804: \begin{equation}
805: H_{\rm out} = H_{\rm big}~.
806: \end{equation}
807: This simplification makes the problem more well-defined; however,
808: every indication is that the characteristic time scale appearing in
809: the youngness paradox is $H_{\rm big}^{-1}$, regardless of this
810: simplification.
811: 
812: 
813: \subsection{The youngness paradox in the square bubble approximation}
814: 
815: We begin with an analysis in the square bubble approximation defined
816: in Sec.~\ref{sec-square}.  By assumption, each bubble appears as a
817: flat patch of the same physical size.  Hence, the comoving volume
818: $V_x$ taken up by a bubble in the outside metric goes like
819: $\exp(-3H_{\rm big} t_{\rm nuc})$.  Note, however, that we have rescaled
820: the Euclidean spatial coordinates inside the bubble, $d{\mathbf y}=\exp(
821:   H_{\rm big} t_{\rm nuc}) d{\mathbf x}$, so as to make the metric in
822: each bubble explicitly the same (not just equivalent by
823: diffeomorphism).  In the sequel, ``comoving volume'' will usually
824: refer to the inside metric and will accordingly be denoted $V_y$.  It
825: will be the same for each bubble, and so will drop out of
826: probabilities.
827: 
828: Let $n(t)$ be the total number of bubbles of our type produced prior
829: to the time $t$.  This grows exponentially with time:
830: \begin{equation}
831:   n(t)=C \exp(3 H_{\rm big}t)~.
832: \label{eq-nour}
833: \end{equation}
834: Here $C$ is a fixed constant, which depends on the size of $\Sigma_0$,
835: the initial state, and the rate at which our vacuum is produced
836: (directly or indirectly) by the fastest inflating vacuum.  This
837: constant will drop out in all ratios.
838: 
839: The nucleation of a bubble like ours will be followed by the formation
840: of observers.  Let $dN^{(1)}$ be the number of observations of some
841: type, in a single bubble of our type, in a comoving volume of size
842: $dV_y$ during the proper time interval $(\tau, \tau+d\tau)$ after the
843: formation of the bubble.  By the homogeneity of the FRW universe,
844: $dN^{(1)}$ will depend only on the FRW time, $\tau$, so we can write
845: \begin{equation}
846:   dN^{(1)} = f(\tau)\, d\tau \, d V_y ~.
847: \end{equation}
848: The function $f(\tau)$ can be thought of as an observer density.
849: 
850: As long as these observations involve looking out into the sky, they
851: will usually be different at different times $\tau$.  For simplicity,
852: we begin by treating $\tau$ itself as an ``observable'', and computing
853: the probability density
854: \begin{equation}
855:   \frac{dp}{d\tau}\propto
856:   \lim_{t\to\infty}\frac{dN}{d\tau}(t)~.
857: \end{equation}
858: This is the probability for an observer to find themselves living a
859: time $\tau$ after the big bang of their bubble.
860: 
861: Both the observer distribution, $f(\tau)$, and the volume per
862: bubble, $V_y$, are the same for all bubbles of our type, by the above
863: assumptions.  Therefore, the total number of $\tau$-observations
864: made by the time $t$ depends on $t$ only through the total number of
865: bubbles $n(t-\tau)$ produced prior to the time $t-\tau$:
866: \begin{eqnarray}
867:   \frac{dN}{d\tau}(t) &=& 
868:   n(t - \tau)   \frac{dN^{(1)}}{d\tau} (\tau)
869:   =
870:   f(\tau)\, V_y\, n(t-\tau) \nonumber \\ 
871:   &=& f(\tau)\, V_y\, \exp \left[3 H_{\rm big} (t-\tau) \right]~.
872: \end{eqnarray}
873: Since the $t$ dependence of the answer is just an overall
874: normalization, it drops out of the probability distribution and we get
875: the simple answer
876: \begin{equation}
877:  \frac{dp}{d\tau}\propto f(\tau) \exp(-3H_{\rm big}\tau) ~.
878:  \label{eq-dpddt}
879: \end{equation}
880: 
881: In our universe, it is reasonable to assume that $f(\tau)$ has a
882: broad (at least Gyr-scale) peak at some $\tau_{\rm peak}\sim O(10$
883: Gyr$)$, since at early times, there was no structure, and at late
884: times, there will be no free energy.  In any case, there will be no
885: features in $f(\tau)$ that can possibly compete with the
886: exponential factor in Eq.~(\ref{eq-dpddt}), which suppresses the
887: probability of late-time observations at a characteristic rate set by
888: the microphysical scale $H_{\rm big}$.
889: 
890: For example, with Planckian $H_{\rm big}\sim O(1)$, there are at any
891: time $t$
892: \begin{equation}
893:   \frac{f(13\, {\rm Gyr})}{f(13.7\, {\rm Gyr})} 
894:   \exp(10^{60})\approx \exp(10^{60})
895: \end{equation}
896: observers who live 13 Gyr after their local big bang, for every
897: observer like us.  Thus, the probability of seeing a 13.7 Gyr old
898: universe with a 2.7 K background temperature is vanishingly small
899: compared to the observation of a warmer CMB and a somewhat younger
900: universe.  
901: 
902: This obviously contradicts experiment.  Note that the probability for
903: what we do see is so small that our observations so far are, by any
904: standards applied in science, perfectly sufficient to rule out the
905: theory---or in this case, the measure.
906: 
907: 
908: \subsection{Explicit conflict with observation}
909: \label{sec-mod}
910: 
911: A possible objection to the above analysis is the fact that $\tau$,
912: the time since the big bang in our bubble, is not a physical
913: observable.  Therefore, following Tegmark~\cite{Teg05}, let us verify
914: explicitly that the youngness paradox manifests itself in the
915: probability distribution for physical observables.
916: 
917: Some observational consequences of the youngness pressure in this
918: measure were described more than ten years ago by Linde, Linde, and
919: Mezhlumian \cite{LinLin96}. There, the authors note that the proper
920: time measure predicts that we are living at the center of an
921: underdense region they refer to as an ``infloid''. The effect they
922: discuss arises because regions which spend less time in slow roll
923: inflation (hence regions which reheat sooner and therefore are
924: underdense) are rewarded.  We focus here on a different effect which
925: is more clearly in conflict with observation: the fact that typical
926: observers see a different temperature than we do.
927: 
928: The probability distribution for the temperature is
929: \begin{equation}
930: {dp \over dT} = \int d\tau {dp \over d \tau} g(T |\tau)~,
931: \end{equation}
932: where $g(T |\tau)$ is the probability distribution for temperatures at
933: a fixed FRW time.  For temperatures not too far from the average
934: value,
935: \begin{equation}
936:   g(T |\tau) \propto {1 \over T_{\rm av}(\tau)} \exp 
937:   \left[
938:     - 10^{10} \left({T - T_{\rm av}(\tau) \over T_{\rm av}(\tau)} \right)^2
939:   \right]~,
940: \end{equation}
941: where $T_{\rm av}(\tau)$ is the average temperature at time $\tau$,
942: and the factor of $10^{10}$ arises due to the magnitude of the density
943: perturbations.  The probability distribution becomes
944: \begin{equation} {dp \over dT} \propto \int d\tau 
945: {f(\tau)  \over T_{\rm av}(\tau)}
946: \exp\left[-3
947:     H_{\rm big} \tau - 10^{10} \left({T - T_{\rm av}(\tau) \over
948:         T_{\rm av}(\tau)} \right)^2 \right]~.
949: \end{equation}
950: 
951: For the moment, let us ignore observations occurring before 10 Gyr,
952: because for early enough times these formulas will break down. For the
953: times under consideration, the average temperature satisfies
954: \begin{equation} {T_{\rm av}(\tau_1) \over T_{\rm av}(\tau_2)} =
955:   \left({\tau_2 \over \tau_1}\right)^{2/3}~.
956: \end{equation}
957: The probability distribution for temperature becomes
958: \begin{widetext}
959: \begin{equation}
960: {dp \over dT}  \propto 
961: \int_{10 ~{\rm Gyr}} d\tau\, \tau^{2/3} f(\tau)  
962: \exp \left[ - 3 H_{\rm big} \tau 
963: - 10^{10}\left( \left(T \over 3.3 K \right) \left(\tau \over 10
964: ~{\rm Gyr}\right)^{2/3} - 1 \right)^2 \right]~.
965: \end{equation}
966: \end{widetext}
967: The dominant factor in the integrand is $\exp(- 3 H_{\rm big} \tau )$,
968: since this factor varies over the microphysical time scale $H_{\rm
969:   big}$, while the other factors vary on much larger time scales.
970: Thus the integral is dominated by the lower limit.  So the probability
971: distribution for the temperature, once fluctuations are taken into
972: account, is just equal to the distribution at the early time cutoff.
973: Dropping a $T$-independent normalization factor, we find
974: \begin{equation}
975: {dp \over d T} \propto g(T |\tau = 10 ~{\rm Gyr}) \propto \exp \left[
976:  - 10^{10} \left({T - 3.3 K \over 3.3 K} \right)^2
977:    \right]~.
978: \label{eq-uf}
979: \end{equation}
980: It is easy to see that this prediction is ruled out at great
981: confidence by our observation that $T=2.7$ K.  The conflict only
982: becomes worse as the early time cutoff is reduced.
983: 
984: 
985: \subsection{Exact treatment of the bubble geometry}
986: \label{sec-bubble}
987: In this subsection, we will improve on the above analysis by taking
988: into account the actual dynamics and shape of bubble walls.  Our
989: treatment will clarify the extent to which the square-bubble
990: approximation is justified, and confirm that the youngness paradox
991: arises in the proper-time measure.
992: 
993: Inside a single bubble, as before we define a function $f(\tau)$ giving
994: the number of observations per unit comoving volume per proper time,
995: \begin{equation}
996:   dN^{(1)} = f(\tau)\, d\tau\, d V_c 
997:   = f(\tau)\, 4 \pi \sinh^2\xi ~d\xi\, d\tau ~.
998: \end{equation}
999: To get the total number of observations, $dN$, at given $(\tau, \xi)$,
1000: we must sum over all bubbles.  We can organize this sum in terms of
1001: the time $t_{\rm nuc}<t$ when each bubble was nucleated. Note that
1002: given the coordinates $(\tau, \xi)$ inside the bubble, there is an
1003: upper limit $ t_{\rm nuc}^{\rm max}(\tau, \xi)$ on the nucleation time
1004: so that the region of interest can be produced before the time
1005: $t$. This relationship was derived in Sec.~(\ref{sec-exact}). The sum
1006: becomes
1007: \begin{equation}
1008:   {dN \over d\tau d\xi}=\int_{0}^{ t_{\rm nuc}^{\rm max}} d t_{\rm nuc}
1009:   \left.\frac{dn}{dt}\right|_{t_{\rm nuc}} {dN^{(1)} \over d\tau d\xi}~,
1010: \label{eq-gg}
1011: \end{equation}
1012: where the bubble production rate $dn/dt$ is still given by
1013: Eq.~(\ref{eq-nour}).  Plugging in, we get
1014: \begin{equation} {dN \over d\tau d\xi} = \int_{0}^{ t_{\rm nuc}^{\rm
1015:       max}} d t_{\rm nuc}\, C \exp \left(3 H_{\rm big} t_{\rm nuc}
1016:   \right) f(\tau)\, 4 \pi \sinh^2\xi
1017: \end{equation}
1018: where, as derived in (\ref{eq-app}),
1019: \begin{equation} 
1020:   t_{\rm nuc}^{\rm max} = t - \tau-H_{\rm
1021:     big}^{-1}\left( \xi + e^{- \xi} - 1 + ... \right)~.
1022: \end{equation}
1023: Performing the integral and dropping constant factors, we get
1024: \begin{eqnarray}
1025:   & &{dN \over d\tau d\xi} = \left[
1026:     e^{3 H_{\rm big} t^{\rm max}_{\rm nuc}} -1 \right]
1027:   ~ f(\tau) \sinh^2 \xi \\ \nonumber 
1028:   &=&\left[e^{3 H_{\rm big} 
1029:       \left[t - \tau -H_{\rm big}^{-1}(\xi +e^{-\xi} - 1 + ...)\right] 
1030:    }- 1 \right]
1031:   f(\tau)\, 
1032:   \sinh^2\xi 
1033: \end{eqnarray}
1034: Taking the limit $t \to \infty$, we can ignore the ``$-1$'' coming
1035: from the lower limit of integration; in this limit the $t$ dependence
1036: is only an overall multiplicative factor which vanishes upon
1037: normalization. Thus we obtain a simple formula for the probability
1038: distribution
1039: \begin{equation}
1040: {d N \over d \xi d \tau} = f(\tau)  e^{-3 H_{\rm big} \tau}~ 
1041:   \sinh^2\xi e^{-3 (\xi+e^{-\xi} - 1 + ...)}
1042: \end{equation}
1043: The striking feature of this probability distribution is that it
1044: factorizes into a function of the spatial coordinate $\xi$ times a
1045: function of the FRW time $\tau$. This is exactly true, because the
1046: ``$\ldots$'' appearing in the formula is a function of $\xi$ only.
1047: 
1048: The distribution as a function of $\tau$ is given by
1049: \begin{equation}
1050: \frac{dp}{d\tau}\propto f(\tau) e^{-3 H_{\rm big} \tau}~.
1051: \label{eq-eyp}
1052: \end{equation}
1053: This distribution is exactly the same as Eq.~(\ref{eq-dpddt}), which
1054: was derived in the square bubble approximation.  So the youngness
1055: paradox appears in exactly the same way in the true geometry.
1056: 
1057: The spatial distribution of observers at a fixed FRW time $\tau$ is
1058: \begin{equation}
1059: \frac{dp}{d\xi}\propto \sinh^2\xi \exp\left[-3 (\xi+e^{-\xi} - 1 +
1060:   ...)\right]
1061: \end{equation}
1062: This distribution peaks at $\xi$ of order one, and falls exponentially
1063: for large $\xi$. So most observers live within a few curvature radii
1064: of the ``center of the universe.''  The center is defined by the
1065: geodesic piercing the bubble nucleation point.
1066: 
1067: \subsection{Why did the square bubble approximation work?}
1068: \label{sec-location}
1069:  The main effect of the correctly computed probability distribution 
1070: over $\xi$ is to allow only an effective comoving volume
1071: \begin{equation}
1072:   V_{\xi, {\rm eff}}=4\pi\int_0^\infty d\xi
1073:   \sinh^2\xi \exp\left[-3(\xi-1+e^{-\xi})\right]\approx 15.75
1074: \end{equation}
1075: to contribute for every bubble.
1076: 
1077: The above result could not have been computed in the square bubble 
1078: approximation, but it explains why that approximation worked for 
1079: computing the {\em temporal\/} distribution of observers.  The point 
1080: is that in a large regime, it is possible to identify the {\em finite\/} 
1081: spatial region containing typical observers with the finite flat patch 
1082: of ``new vacuum'' inserted by hand in the square bubble approximation.  
1083: 
1084: This identification is not a true match, because of the different
1085: spatial curvature. But during and after inflation, there is a long
1086: period where curvature is negligible and the scale factor would be the
1087: same function of time in a flat FRW universe.  If the observations
1088: contributing to the measure occur in this regime, then the use of
1089: spatially flat time-slices in the square-bubble approximation will be
1090: legitimate.
1091: 
1092: The effective physical volume at large FRW time $\tau$ is $15.75\,
1093: a^3(\tau)$.  During inflation, $a(\tau)=H_{\rm inf}^{-1}\sinh H_{\rm
1094:   inf}\tau$.  To match this to the square bubble physical volume, $V_y
1095: H_{\rm inf}^{-3} \exp\left(3H_{\rm inf} \tau \right)$, at $\tau \gg
1096: H_{\rm inf}^{-1}$, requires the choice
1097: \begin{equation}
1098:   V_y=V_{\xi, {\rm eff}}/8\approx 2~.  
1099: \label{eq-vy}
1100: \end{equation}
1101: 
1102: Note that in a problem involving different types of bubbles, the
1103: physical volume of a new bubble will be of order $H_{\rm inf}^{-3}$.
1104: Generically $H_{\rm inf}$ will be smaller than the outside Hubble
1105: constant.  If we took Eq.~(\ref{eq-vy}) literally, the square bubble
1106: approximation would involve replacing a large number of outside Hubble
1107: volumes with the new vacuum.  This contradicts the geometric fact that
1108: asymptotically, the new bubble takes up the comoving volume occupied
1109: by only one outside Hubble volume at the time of nucleation.  Of
1110: course, the choice of $V_y$ dropped out of ratios, so it could be
1111: reduced without affecting relative probabilities.
1112: 
1113: In any case, while the square bubble approximation turned out to be a
1114: useful shortcut under the above assumptions, it is just as simple, and
1115: much more reliable, to use the exact geometry, as encoded in
1116: Eq.~(\ref{eq-exact}), to compute probabilities.
1117: 
1118: 
1119: 
1120: \section{Modifications of the proper time measure}
1121: \label{sec-fixes}
1122: 
1123: Obviously, the result that practically all observers live at a much
1124: earlier time, and see a very different universe, than we do, is fatal
1125: for the proper time measure.  Perhaps the measure can be modified in
1126: some way, so as to avoid this problem? 
1127: 
1128: \subsection{Don't ask, don't tell}
1129: 
1130: Linde advocates a simple resolution to the youngness paradox in
1131: Ref.~\cite{Lin06} (see also references therein).  One should simply
1132: not ask how long after reheating the typical observers form, but
1133: merely compute the rate at which reheating hypersurfaces of different
1134: inflating vacua are produced.  This restriction has a number of
1135: problems.  If we cannot ask about the temperature measured by a
1136: typical observer, the measure is not complete.  Moreover, if we cannot
1137: ask about observers, then we cannot count them, and so we cannot
1138: condition on their number.  This would eliminate the anthropic
1139: solution to the cosmological constant problem.  And finally, as noted
1140: in Ref.~\cite{Lin07}, this restriction does not fully solve the
1141: youngness problem in any case.  It merely confines the problem to
1142: effects before reheating.  In particular, it gives overwhelming weight
1143: to vacua with a shorter period of inflation, and thus predicts a wide
1144: open universe.  Thus, a different modification is needed.
1145: 
1146: A general idea for such a fix was outlined in Ref.~\cite{Lin07}: ``One
1147: should compare apples to apples, instead of comparing apples to the
1148: trunks of the trees.''  In other words, we should assign a correction
1149: factor $e^{3 H_{\rm big} \Delta t_i}$ to the probability $p_i$ for the
1150: observation $O_i$, where $\Delta t_i$ is the amount of time it takes
1151: to produce such an observation, in some relative sense to be defined
1152: below.  The corrected relative probabilities are thus:
1153: \begin{eqnarray}
1154:   \frac{P(O_1)}{P(O_2)} &=& 
1155:   \frac{p(O_1)}{p(O_2)} \exp[3 H_{\rm big} (\Delta t_1-\Delta t_2)]
1156:   \\ \nonumber
1157:   &=&\lim_{t\to\infty}\frac{N_1(t)}{N_2(t)} 
1158:   \exp[3 H_{\rm big} (\Delta t_1-\Delta t_2)]~.
1159: \label{eq-fix}
1160: \end{eqnarray}
1161: Compared to Eq.~(\ref{eq-prob}), this bolsters mature folk like
1162: ourselves, by compensating for the enormous volume growth that
1163: Boltzmann babies can take advantage of.
1164: 
1165: No general, sharp definition of $\Delta t_i$ was attempted in
1166: Ref.~\cite{Lin07}, where explicit calculations were carried out only
1167: for a model containing two vacua with different lengths of inflation;
1168: $\Delta t_i$ was defined to be the duration of inflation in each
1169: vacuum.  While Ref.~\cite{Lin07} claimed that the procedure also
1170: resolves other aspects of the youngness paradox, such as the
1171: overwhelming probability for a hotter universe, it offered no
1172: definition of $\Delta t$ in that context, nor did it display an
1173: explicit computation of the corrected probability.
1174: 
1175: In fact, we have been unable to come up with a general definition of
1176: $\Delta t$ that succeeds in fixing the youngness problem.  This does
1177: not mean that it cannot be done.  But perhaps it will help sharpen the
1178: challenge if we discuss a few proposals that may come to
1179: mind.\footnote{We thank Andrei Linde for discussions that influenced
1180:   some of the definitions explored below.  However, we make no claim
1181:   that any of them reflect his views accurately.}
1182: 
1183: \subsection{Spatial averaging}
1184: 
1185: The prediction that we should observe a warmer CMB temperature arose
1186: from the fact that it takes longer to produce observers who see low
1187: CMB temperatures.  To get a more reasonable probability distribution
1188: for the CMB temperature, we want to eliminate the enormous cost of
1189: waiting for the universe to cool off.  It seems reasonable to assign
1190: $\Delta t(T)$ as the amount of FRW time until the average background
1191: temperature is $T$. Since we will only be comparing observations
1192: within the same bubble and after reheating, an additive constant in
1193: $\Delta t$ is unimportant and so we can start our clock at any time we
1194: like.
1195: 
1196: We explore this proposal mostly because it seems like the most
1197: straightforward and naive fix.  In fact, it is unclear how the above
1198: definition would generalize to observables that can take on the same
1199: value at very different times.  Even for the temperature, small
1200: perturbations render the relation between its average value and a
1201: particular time slice ambiguous at the $10^{-5}$ level---much too
1202: large to define $\Delta t$ with the required Planckian precision.  We
1203: will disregard all these issues, since the modification fails even in
1204: the idealized special case we consider.
1205: 
1206: For temperatures and times close to those we observe, the average
1207: temperature on $\tau=$ const slices satisfies
1208: \begin{equation} {T(\tau_1) \over T(\tau_2)} = \left({\tau_2 \over
1209:       \tau_1}\right)^{2/3}~.
1210: \end{equation}
1211: Thus, $\Delta t(T)$ is given by
1212: \begin{equation}
1213:   \Delta t(T) = (13.7~ {\rm Gyr}) 
1214:   \left( 2.7~{\rm K} \over T \right)^{3/2}~.
1215: \label{eq-classical}
1216: \end{equation}
1217: 
1218: The modification fails because it is comparatively easy to find
1219: deviations from the average temperature.  To see this, let us begin by
1220: considering a further idealization: Let us exclude fluctuations of
1221: $T$.  In other words, we will assume that the CMB temperature, at
1222: fixed $\tau$, is given everywhere precisely by the same value.  
1223: 
1224: With this additional idealization, the modification actually works!
1225: There is now a one-to-one correspondence between $\tau$ and $T$, so we
1226: can use Eq.~(\ref{eq-dpddt}) to obtain the (unmodified) probability
1227: distribution for $T$:
1228: \begin{equation} {dp \over dT} = {dp \over d\tau} {d\tau \over dT}
1229:   \propto f(\tau(T)) {d\tau \over dT} \exp \left[-3 H_{\rm big}
1230:     \tau(T) \right]
1231: \label{eq-cluf}
1232: \end{equation}
1233: where $f(\tau)$, as before, is the rate of observations per comoving
1234: volume per unit time per bubble. The quantity $f(\tau(T)) {d\tau \over
1235:   dT} $ is naturally identified as $f(T)$, the rate of observations
1236: per comoving volume per unit background temperature per bubble.  Using
1237: the previous formula relating time to temperature, we find
1238: \begin{equation} {dp \over dT} \propto f(T) \exp\left[- 3 H_{\rm big}
1239:     \cdot (13.7~{\rm Gyr}) \left(2.7~{\rm K} \over T\right)^{3/2}
1240:   \right] ~.
1241: \end{equation}
1242: 
1243: Still working in the idealization of exactly homogeneous background
1244: temperature, let us now compute the modified probability distribution
1245: for temperature. It is
1246: \begin{equation} {dP \over dT} \propto f(T) \exp \left[-3 H_{\rm big}
1247:     (\tau(T) - \Delta t(T)) \right]
1248: \end{equation}
1249: We have defined $\Delta t$ so that the exponent is zero, so the modified 
1250: probability distribution for temperature is simply proportional to the 
1251: number of observations at each temperature,
1252: \begin{equation}
1253: {dP \over dT}  \propto f(T)~.
1254: \label{eq-dream}
1255: \end{equation}
1256: This answer seems intuitive and has no youngness problem. (See,
1257: however, the discussion at the end of Sec.~\ref{sec-anticipate}.)
1258: 
1259: Once we allow for fluctuations of the temperature, $\Delta t(T)$ can
1260: still be defined in terms of the average temperature.  But our recipe
1261: for repairing the probabilities no longer works.
1262: 
1263: Now the starting point is the (unmodified) probability distribution
1264: obtained in Sec.~\ref{sec-mod}, Eq.~(\ref{eq-uf}).  After applying
1265: Eq.~(\ref{eq-fix}), with $\Delta t(T)$ given by
1266: Eq.~(\ref{eq-classical}), we obtain the modified distribution
1267: % \begin{equation} {dp \over dT} \propto \exp \left[3 H_{\rm big}\Delta
1268: %     t(T)\right] \int_{10 ~{\rm Gyr}} d\tau \exp \left[ - 3 \cdot
1269: %     10^{61} \left( \tau \over 10 ~{\rm Gyr} \right) - 10^{10}\left(
1270: %       \left(T \over 3.3 K \right) \left(\tau \over 10 ~{\rm
1271: %           Gyr}\right)^{2/3} - 1 \right)^2 \right]
1272: % \end{equation}
1273: % The integral is unchanged, so it is still dominated by the lower
1274: % limit, and we get
1275: \begin{equation} {dP \over d T} = \exp \left[3 H_{\rm big}\Delta t(T)
1276:   \right] g(T |\tau = 10 ~{\rm Gyr})
1277: \end{equation}
1278: Using 
1279: \begin{equation}
1280:   \Delta t(T) = (10 ~{\rm Gyr}) \left( 3.3 K \over T \right)^{3/2}~,
1281: \end{equation}
1282: and assuming Planckian $H_{\rm big}$, we get
1283: \begin{equation}
1284: {dP \over d T} = \exp \left[ 10^{61}  \left( 3.3 K \over T
1285:   \right)^{3/2}
1286:  - 10^{10} \left({T - 3.3 K \over 3.3 K} \right)^2
1287:    \right]
1288: \end{equation}
1289: 
1290: The temperature is now driven to the {\em lowest\/} possible value.
1291: It is still favorable to live early, and because of the primordial
1292: density fluctuations, it is not all that hard to find an anomalously
1293: cool region even at early times.  Our modification factor rewards us
1294: for this as if we had honestly waited until the average temperature
1295: becomes so low.  Thus, it overcompensates.
1296: 
1297: This new distribution is also ruled out, at enormous confidence level,
1298: by our observation of 2.7 K.
1299: 
1300: \subsection{Waiting for the first time}
1301: 
1302: Another possibility is to define $\Delta t_i$ for the observation of
1303: type $O_i$ as the time it takes the universe, starting from the
1304: beginning of time (the slice $\Sigma_0$), to produce the first such
1305: observation.
1306: 
1307: Thus defined, $\Delta t_i$---and hence, the corrected
1308: probabilities---will depend on the initial conditions on $\Sigma_0$.
1309: This dependence may be mild, and in any case we can see no reason why
1310: probabilities (at least for some observables) should not depend on the
1311: initial conditions of the universe.  However, if we define $\Delta
1312: t_i$ as the time when $N_i$ jumps from 0 to 1, then it will also
1313: depend on accidents of the semiclassical evolution at early times,
1314: such as the time when a particular tunneling event happens to take
1315: place, and we would not be able to compute it directly from the
1316: theory.
1317: 
1318: This problem can be resolved by defining $\Delta t_i$ to be the time
1319: when the expectation value $\langle N_i(\Delta t)\rangle$ becomes
1320: 1.\footnote{More generally, one could consider defining $\Delta t$ to
1321:   be the time when the expectation value reaches some fixed value
1322:   $N_{\rm min}$.  In Eq.~(\ref{eq-vp}) below, the small volume limit
1323:   is equivalent to taking $N_{\rm min}\to\infty$ at fixed $V$.}  This
1324: still depends on initial conditions but can be computed from the
1325: theory in the semiclassical regime.
1326: 
1327: However, this definition conflicts with an important property of
1328: probabilities.  Consider the special case that $O_1$ and $O_2$ are
1329: mutually exclusive outcomes of an experiment.  For example, outcome
1330: $O_1$ ($O_2$) may be up (down) when the spin of a single electron is
1331: measured by a man in a penguin suit.  In general there may be
1332: additional possible outcomes $O_3,\ldots$, but in any case, it must be
1333: true that
1334: \begin{equation}
1335: p_1+p_2=p_{12}~,
1336: \end{equation}
1337: where $p_{12}$ is the probability for the outcome ``1 or 2''.  Indeed,
1338: this property will be satisfied by the original probabilities defined
1339: in Eq.~(\ref{eq-prob}).
1340: 
1341: However, $\Delta t_{12}<\Delta t_i$, $i=1,2$, because the expected
1342: time when ``1 or 2'' is first observed is simply the time when the
1343: experiment is first likely to be performed.  This is sooner than the
1344: expected time when, say, 1 is first observed, since the very first
1345: experiment can only have one outcome.  Therefore, the corrected
1346: probabilities do not add up correctly:
1347: \begin{eqnarray}
1348:   P_1+P_2&=&
1349:   p_1 e^{3 H_{\rm big} \Delta t_1} +p_2 e^{3 H_{\rm big} \Delta t_2}
1350:   \\ \nonumber
1351:   &>&(p_1+p_2) e^{3 H_{\rm big} \Delta  t_{12}} = P_{12}~,
1352: \end{eqnarray}
1353: Another way of saying this is that we can change the total probability
1354: for a set of alternative outcomes by whether we view the alternatives
1355: separately and add probabilities, or group the alternatives together
1356: and directly compute the probability for this compound outcome.  This
1357: is clearly absurd.
1358: 
1359: A particularly simple and striking result obtains if we assume initial
1360: conditions that are already in the stationary regime.  Then
1361: \begin{equation}
1362:   \langle N_i(t)\rangle  = V p_i (e^{3 H_{\rm big} t}-1)~.
1363: \label{eq-nstat}
1364: \end{equation}
1365: The uncorrected probabilities $p_i$ are dynamically determined by the
1366: attractor behavior; the overall scaling $V$ depends on the volume of
1367: $\Sigma_0$.  We find for the correction factor
1368: \begin{equation}
1369: e^{3 H_{\rm big} \Delta t_i}= \frac{Vp_i+1}{Vp_i}~,
1370: \end{equation}
1371: and hence
1372: \begin{equation}
1373: \frac{P_i}{P_j}=\frac{V p_i + 1}{V p_j+1}.
1374: \label{eq-vp}
1375: \end{equation}
1376: In the large volume limit, there is no correction and the youngness
1377: paradox persists.  For finite $V$, the corrected probabilities do not
1378: obey $P_1+P_2=P_{12}$.  In the limit $V\to 0$, all alternatives become
1379: equally likely, $P_i/P_j=1$, no matter how they were defined!
1380: 
1381: 
1382: \subsection{Growing together}
1383: 
1384: A different definition for $\Delta t_i$ may be motivated by another
1385: quote from Ref.~\cite{Lin07}: $\Delta t_i$ is ``the time when the
1386: stationary regime becomes established'' for the observation $O_i$.
1387: Mathematically, we may attempt to capture this idea as follows.  At
1388: late times, we know that the number of observations of any type will
1389: grow as $e^{3 H_{\rm big} t}$, so $\dot N_i/N_i\to 3 H_{\rm big} $.
1390: At any finite time, there will be a small correction to this time
1391: dependence, so we may define $\delta t_i$ as the earliest time when
1392: \begin{equation}
1393:   \left|1-\frac{\dot N_i(t_i)}{3 H_{\rm big} N_i(t_i)}\right|
1394:   \leq \epsilon~.
1395: \end{equation}
1396: 
1397: It would seem arbitrary to specify an particular small deviation
1398: $\epsilon$ beyond which we consider the stationary regime established.
1399: Therefore, let us take the limit $\epsilon\to 0$.  In this limit each
1400: $\delta t_i$ contains the same additive divergence, $-\log \epsilon$,
1401: which we discard and define
1402: \begin{equation}
1403: \Delta t_i=\lim_{\epsilon \to 0}\delta t_i + \log\epsilon~.
1404: \end{equation}
1405: 
1406: To see that this measure does not work, let us focus again on the CMB
1407: temperature in our own vacuum.  For the sake of argument, suppose that
1408: no observers exist prior to some cutoff FRW time, say, $\tau_{\rm
1409:   min}=10$ Gyr.  By the results of Sec.~\ref{sec-mod}, for any finite
1410: geodesic time $t$, practically all observations of {\em any\/} value
1411: of $T$ are made within a time of order $H_{\rm big}^{-1}$ after
1412: $\tau_{\rm min}$.  Therefore, to accuracy $H_{\rm big}^{-1}$,
1413: $|1-{\dot N_T}/(3 H_{\rm big} N_T)|$ will drop below $\epsilon$ at the
1414: same time $t$, for any temperature $T$.  Hence, $\Delta t_T$ is
1415: independent of $T$ to this accuracy.\footnote{A finite width $dT$ is
1416:   implicit; see Eq.~(\ref{eq-density}) and the footnote thereafter.
1417:   We use $N_T$ as a short notation for $\frac{dN}{dT} dT$. Our
1418:   conclusion becomes strictly true only in the $\epsilon \to 0$ limit,
1419:   when the total volume of the FRW cutoff surfaces between $\Sigma_0$
1420:   and $\Sigma_t$ becomes large enough to contain all possible values
1421:   of $T$.}  Therefore, our ``modification'' does not in fact change
1422: relative probabilities at all.
1423: 
1424: To complete this argument, we should now take the cutoff FRW time,
1425: $\tau_{\rm min}$, earlier and earlier, until it is removed altogether.
1426: However, this introduces only information about early universe physics
1427: into our modification of the proper time measure.  It cannot possibly
1428: restore a reasonable probability distribution for the CMB temperature
1429: measured by observers in the present era.
1430: 
1431: It is interesting that like ``spatial averaging'', the present
1432: modification {\em would\/} have worked for (fictitious) observables
1433: that are in one-to-one correspondence with the FRW time.  In this
1434: case, it would not have been true that at any time $\tau$, there is a
1435: nonzero amplitude for any temperature $T$.  Instead,
1436: \begin{equation}
1437:   N_{T_1}(t)=
1438:   \frac{f(T_1)}{f(T_2)}N_{T_2}(t+\tau(T_2)-\tau(T_1))~,
1439: \end{equation} 
1440: and hence,
1441: \begin{equation}
1442:   \frac{\dot N_{T_1}}{N_{T_1}} (t)= 
1443:   \frac{\dot N_{T_2}}{N_{T_2}}(t+\tau(T_2)-\tau(T_1))~.
1444: \end{equation}
1445: Then we would have found
1446: \begin{equation}
1447: \Delta t_1-\Delta t_2=\Delta t(T_1)-\Delta t(T_2)~,
1448: \end{equation}
1449: and we would have recovered the intuitive result of
1450: Eq.~(\ref{eq-dream}).
1451: 
1452: Apparently, the problem with both of these modifications is that the
1453: {\em value\/} of an observable does not give us enough information
1454: about the FRW {\em time\/} when the observation is made---but this is
1455: precisely the time we would like to use for $\Delta t$.  This
1456: motivates our final attempt at modifying the proper time measure, in
1457: which $\Delta t$ is defined not as a function of observables, but
1458: directly as a function of the time when the observation is made,
1459: regardless of its outcome.
1460: 
1461: 
1462: \subsection{Anticipation}
1463: \label{sec-anticipate}
1464: 
1465: Instead of tying $\Delta t$ to a specific observable, we can go back
1466: and fix the time shift directly for the geodesics.  Effectively, it is
1467: like ``anticipating'' all observations that will happen in a given
1468: bubble.  More precisely, let us project every observation $O_i$ inside
1469: a bubble back to the most recent bubble wall along the geodesics of
1470: the congruence, and count it toward $N_i(t)$ as soon as the relevant
1471: portion of the wall lies below $\Sigma_t$.  This amounts to choosing
1472: $\Delta t_i$ to be the geodesic time between the domain wall and the
1473: observation.  It is not difficult to see that this choice eliminates
1474: any pressure to make observations very early, and that it reduces to
1475: the $\Delta t$'s used in the specific example of
1476: Ref.~\cite{Lin07}. However, in general it suffers from two major
1477: problems.
1478: 
1479: First, it is not sharply enough defined.  The prescription involves
1480: projecting onto domain walls.  These objects have an inherent
1481: thickness, which can be microscopic, but need not be Planckian. This
1482: is a problem because we need a proposal which is well-defined at the
1483: length scale $H_{\rm big}^{-1}$, which may be Planckian.  Moreover, an
1484: approximately defined object like a domain wall has no place in a
1485: fundamental definition of probabilities for all observations.  There
1486: is a smooth interpolation between objects that appear obviously
1487: recognizable as domain walls, and general field configurations.  (This
1488: objection could be raised also against other measures that involve
1489: domain walls in their definition, such as Ref.~\cite{GarSch05}.)
1490: 
1491: On the observational side, the projection method suffers from the
1492: ``Boltzmann brain'' problem~\cite{Pag06,BouFre06b}. The reason is that we
1493: are now completely indifferent to when observations inside the bubble
1494: are made. By the results of Sec.~\ref{sec-location}, we may focus on a
1495: single comoving volume at the center of any metastable de~Sitter
1496: bubble with sufficiently small cosmological constant (such as,
1497: presumably, our own vacuum).  An infinite number of observers are
1498: formed at late times in this volume, due to rare thermal
1499: fluctuations~\cite{Pag06}.  All of these Boltzmann brains will be
1500: projected back, and so will dominate over other observers.  Thus, with
1501: probability 1, we should be Boltzmann brains, which is in conflict
1502: with observation~\cite{DysKle02}.  (Alternatively, we can interpret
1503: this infinity as telling us that projection defeats the most basic
1504: purpose of the measure, which is to regulate the infinities occurring
1505: in eternal inflation.)
1506: 
1507: Mathematically, the Boltzmann brain problem shows up as follows.  The
1508: effect of the ``anticipation'' modification is to render the
1509: temperature distribution apparently well-behaved: we have finally
1510: succeeded in producing the hoped-for Eq.~(\ref{eq-dream}).  But this
1511: equation is a poisoned chalice: it is not as harmless at it looks.
1512: Boltzmann brains arise at a fixed rate per unit time and unit physical
1513: volume, not per comoving volume.  Thus, the comoving observer density
1514: $f(\tau)$ grows exponentially with the scale factor at extremely late
1515: times, so $f(T)$ diverges at the Hawking temperature of the de~Sitter
1516: space.
1517: 
1518: \acknowledgments We are grateful to Andrei Linde for extensive
1519: discussions.  This work was supported by the Berkeley Center for
1520: Theoretical Physics, by a CAREER grant of the National Science
1521: Foundation, and by DOE grant DE-AC03-76SF00098.
1522: 
1523: \bibliographystyle{board}
1524: \bibliography{all}
1525: 
1526: \end{document}
1527: