hep-th0304199/npet
1: %\documentclass[12pt]{article}
2: \documentstyle[12pt,epsf]{article}
3: \setlength{\topmargin}{-.3in}
4: \setlength{\oddsidemargin}{.0in}
5: %\setlength{\evensidemargin}{-.21in}
6: \setlength{\textheight}{8.2in}
7: \setlength{\textwidth}{6.0in}
8: \arraycolsep=2pt
9: \renewcommand{\theequation}{\thesection.\arabic{equation}}
10: \def\be{\begin{equation}}
11: \def\ee{\end{equation}}
12: \def\beq{\begin{eqnarray}}
13: \def\eeq{\end{eqnarray}}
14: \def\e{\epsilon}
15: \def\l{\lambda}
16: \def\g{\gamma}
17: \def\p{\partial}
18: \def\t{\tilde}
19: \def\G{\Gamma}
20: \def\R{{\cal R}}
21: \def\RR{{\bf R^3}}
22: \def\({\left (}
23: \def\){\right )}
24: \def\[{\left [}
25: \def\[{\right ]}
26: \begin{document}
27: 
28: \begin{titlepage}
29: \bigskip
30: \rightline{}
31: \rightline{hep-th/0304199}
32: \bigskip\bigskip\bigskip\bigskip
33: \centerline{\Large \bf {Negative Energy Density in Calabi-Yau
34: Compactifications}}
35:      \bigskip\bigskip
36:       \bigskip\bigskip
37: 
38:   \centerline{\large Thomas Hertog${}^1$, Gary T. Horowitz${}^1$,
39:   and Kengo Maeda${}^2$}
40:      \bigskip\bigskip
41:   \centerline{\em ${}^1$ Department of Physics, UCSB, Santa Barbara, CA 93106}
42:   \centerline{hertog@vulcan.physics.ucsb.edu, 
43:    gary@physics.ucsb.edu }
44: 	    \bigskip
45:  \centerline{\em ${}^2$ Yukawa Institute for Theoretical Physics,
46:   Kyoto University, Kyoto 606-8502, Japan}	    
47: 	    \centerline{kmaeda@yukawa.kyoto-u.ac.jp}
48: 	    \bigskip\bigskip
49: 
50: \begin{abstract}
51: 
52: We show that a large class of supersymmetric compactifications, including all
53: simply connected Calabi-Yau and $G_2$
54: manifolds, have classical configurations with negative energy density as
55: seen from four dimensions. In fact, the energy density can be
56: arbitrarily negative -- it is unbounded from below. Nevertheless, 
57: positive energy theorems show that the
58: total ADM energy remains positive. Physical consequences of the
59: negative 
60: energy density include new thermal instabilities, and possible violations
61: of cosmic censorship.
62: 
63: \end{abstract}
64: \end{titlepage}
65: 
66: \baselineskip=18pt
67: 
68: 
69: \setcounter{equation}{0}
70: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
71:  \section{Introduction}
72: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
73:  
74: Standard supersymmetric compactifications of string theory
75: consist of solutions of the form $M_4\times K$ where $M_4$ is
76: four dimensional Minkowski spacetime and $K$ is a compact, Ricci flat
77: manifold admitting a covariantly constant spinor. Familiar examples of $K$
78: include $T^n$, $K3$, Calabi-Yau spaces, and manifolds with $G_2$ holonomy.
79: It turns out that there is an important
80: qualitative difference between some of these supersymmetric
81: compactifications and others. We will show that a large class, 
82: including all simply connected
83: Calabi-Yau and $G_2$ manifolds, have  the surprising property
84: that there are configurations with negative energy
85: density. In other words, from a four dimensional perspective, there can
86: be finite regions of space with negative energy. In fact, the energy
87: density is unbounded from below! In contrast, these
88: properties do not hold - at least not in the same way - 
89: for $T^n$ or $K3$ compactifications.
90: 
91: We will work in the context of classical vacuum solutions to higher
92: dimensional general relativity.  So there is no matter or energy density in
93: the higher dimensional space. The negative energy arises from the Kaluza-Klein
94: compactification. These results clearly apply to string theory since
95: vacuum solutions are all (approximate) solutions to string
96: theory and M theory with the 
97: other fields set to zero. We should emphasize that the negative energy
98: regions we have in mind are very different from previous discussions
99: of negative energy in Kaluza-Klein theory \cite{Witten:gj,Brill:di}. 
100: In those cases,
101: supersymmetry was broken and the spacetime was not topologically a product
102: ${\bf R}^4\times K$. In contrast, the
103: compactifications we consider here are all supersymmetric and the spacetime
104: will be a product manifold.
105: 
106: The key mathematical fact which allows configurations of negative energy density
107: is the following. As we review in section 2, all simply connected
108: compact manifolds of dimension five, six, or seven admit 
109: Riemannian metrics with
110: positive scalar curvature. Other manifolds, such as $T^n$ and $K3$ do not.
111: Positive scalar curvature on $K$ leads to negative energy density as follows.
112: Vacuum solutions can
113: be characterized by their initial data on $\RR \times K$.
114:  Since we want to minimize the energy, we
115: set the time derivatives of the metric to zero. For time symmetric
116: initial data, the Einstein constraint equations reduce to the vanishing
117: of the scalar curvature, $\R=0$.
118: For a product metric on $\RR \times K$, $\R = \R_3 + \R_K$ where $\R_3$ is
119: the scalar curvature on $\RR$ and $\R_K$ is the scalar curvature on $K$.
120: If $\R_K >0$, we must take $\R_3 <0$. But negative
121: scalar curvature on $\RR$ is just like negative energy density. (Recall that
122: the usual constraint of $3+1$ dimensional general relativity says\footnote{We
123: set $8\pi G =1$.}
124: $\R_3 = 2\rho$
125: in the time symmetric case.) Therefore, from an effective four dimensional
126: standpoint, positive scalar curvature on $K$ acts 
127: like negative energy density. In other words, ten dimensional
128: vacuum gravity has configurations with effective 
129: negative energy density! Of course, we must require that the metric on $K$ 
130: approaches the standard Ricci flat metric at infinity, so we cannot
131: keep the metric a product everywhere.  However, one can satisfy this
132: boundary condition and keep the region of negative energy density by
133: taking the  metric to be product inside
134: a large ball of radius $R_0$. In a finite transition region, $R_0 < r < R_1$,
135: one can change the metric on $K$ to the standard Ricci flat metric.
136: 
137: Not only is there negative energy density in four dimensions, but this
138: energy can be arbitrarily negative. This follows immediately from the
139: fact that there is no upper bound on the scalar curvature $\R_K$.
140: Given a  metric on $K$ with $\R_K>0$, one can clearly rescale it by
141: a constant factor and make the scalar curvature arbitrarily large.
142: This shows that the negative energy density is unbounded from below.
143: %*********************
144: (Of course, once the curvature becomes larger than the string scale, there
145: may be significant $\alpha'$ corrections.)
146: 
147: As we discuss in section 2, the metrics with positive scalar curvature on
148: $K$ lie a finite distance away from the moduli space of Ricci flat metrics.
149: This shows that the negative energy is a nonperturbative effect, which
150: cannot be seen at any finite order in perturbation theory about the moduli 
151: space. 
152: In terms of an effective four dimensional potential, we will see that
153: (part of) the negative energy density can be described by a scalar
154: field with a negative potential which falls off exponentially. Negative
155: potentials are familiar in supersymmetric
156: AdS compactifications, but to our knowledge, this is the first time
157: they have been seen in supersymmetric asymptotically flat compactifications.
158: 
159: Given the existence of negative energy density, it is natural to ask if the
160: total ADM energy can be negative. Recall that
161: the total energy for any spacetime which asymptotically 
162: approaches $M_4\times K$ is well-defined. In the construction outlined above, 
163: the negative energy grows like the volume of the ball. The transition region
164: includes positive energy, but one might expect this only grows like the area 
165: so the total energy could be negative. This would imply that
166: these supersymmetric compactifications are unstable.  However, this
167: does not occur. There are positive energy theorems 
168: \cite{Witten:mf,Boucher:1984yx,Dai} 
169: which ensure that the total energy remains positive. Coleman and 
170: De Luccia \cite{Coleman:1980aw}
171: showed long ago that gravity can stabilize a false vacuum. (This was
172: applied to supersymmetric vacua in  \cite{Weinberg:id,Cvetic:1992st}.)
173: They considered a scalar field with a potential with two local 
174: minima that are close together. We have discovered 
175: an extreme generalization of this phenomenon: gravity
176: is stabilizing a false vacuum in a theory with a
177: potential which is unbounded from below. This is a flat space analog of the
178: familiar situation for supersymmetric AdS vacua 
179: \cite{Breitenlohner:jf,Gibbons:aq}.
180: 
181: Despite the fact that the total energy remains positive,
182: the existence of regions of negative energy does have physical
183: consequences which we begin to explore. These
184: include new instabilities at finite temperature, and possible violations
185: of cosmic censorship.
186: 
187: An outline of this paper is as follows. In section 2 we review the mathematical
188: results about the existence of metrics of positive scalar curvature that
189: we will need. The next section shows how to use these results to
190: construct configurations with negative energy density. It also
191: gives a qualitative
192: discussion of the effective four dimensional potential that results from 
193: compactification. To gain intuition for how the total energy can remain
194: positive given the unbounded negative potential, we discuss a simple
195: model of four dimensional gravity coupled to a single scalar field
196: in section 4. The following section reviews the positive energy theorems which
197: guarantee that the ADM energy is never negative. In section 6 we discuss
198: some of the physical consequences of the negative energy density. The
199: final section contains remarks about other applications of our results,
200: e.g., to the stability of de Sitter spacetime, and using 
201: S-duality to obtain negative energy density even in K3 compactifications.
202: 
203: 
204: \setcounter{equation}{0}
205: \section{Metrics with positive scalar curvature}
206: 
207: Since metrics with positive scalar curvature play such an important role
208: in our construction, we review here some of the main mathematical results
209: that we will need. (For a more complete discussion of results up to 1989,
210: see \cite{Lawson89}.) Let $K$ be a compact manifold.
211: At first sight, finding a metric whose scalar
212: curvature is positive at every point on $K$ sounds easy, since it is only one
213: scalar inequality on the entire metric. Indeed, every compact $K$ (of
214: dimension $n>2$) admits metrics
215: with negative scalar curvature. However it is known that there is a
216: topological obstruction to the existence of metrics with  positive scalar
217: curvature. For example, the torus $T^n$ (for any $n$) does
218: not admit any metric with\footnote{In this section we will only consider
219: the scalar curvature of $K$ and drop the subscript $K$ on $\R$.} 
220: $\R>0$ \cite{Schoen79}. The same is true for $K3$.
221: In this case the proof is easy: 
222: The index for the Dirac operator on $K3$ is
223: nonzero. This means that for every metric on $K3$, there are nonzero
224: spinors which solve the Dirac equation: $\gamma^i \nabla_i \e =0$.
225: However, if we square the Dirac
226: operator we get
227: \beq
228: \nabla^2 \e - {1\over 4}\R \e =0. 
229: \eeq
230: Multiplying by the adjoint spinor and integrating over $K$ yields
231: \beq
232: \int |\nabla \e|^2 + {1\over 4}\R |\e|^2 =0. 
233: \eeq
234: This shows that the metric cannot have positive scalar curvature. This
235: argument clearly applies to any manifold for which the index of the Dirac
236: operator, $\hat A(K)$, is nonzero.
237: 
238: There is a generalization of $\hat A(K)$ which
239: has been shown to completely characterize when a (simply connected, spin)
240: manifold admits a metric
241: of positive scalar curvature. It is usually denoted $\alpha(K)$, and was
242: first introduced in 1974 by Hitchen \cite{Hitchen74}
243: who showed that if there is a metric 
244: with $\R>0$, then $\alpha(K) =0$. The converse was established 
245: by Stolz in 1990 \cite{Stolz90}:
246: If $\alpha(K) =0$, there always exists a metric with
247: $\R>0$.
248: If the dimension of the manifold is a
249: multiple of four,  $\alpha(K)$ is proportional to $\hat A(K)$.
250: In general, the definition of $\alpha(K)$ is more involved, but for
251: our purposes, its most important property is that it can be nonzero
252: only in dimensions $n= 0,1,2,4\ {\rm mod}\ 8$. (This assumes $n>4$.)
253: So there is a possible obstruction to positive scalar curvature
254: only in these dimensions. Even in these dimensions,
255: this obstruction only applies
256: to spin manifolds: If $K$ does not admit spinors, then it always admits a metric
257: of positive scalar curvature \cite{Lawson80}.
258: Taken together, these results show the following: 
259: 
260: {\it Every simply connected manifold of dimension 5, 6, or 7 admits
261: a metric with
262: positive scalar curvature.}
263: 
264: The situation for nonsimply connected manifolds is much more complicated
265: and still not well understood \cite{Dwyer02}.
266: We have already remarked that
267: the torus does not admit a metric with $\R>0$ even in $5,6,7$ dimensions.
268: However, it has been shown that if $K$ is simply connected and admits a metric
269: of positive scalar curvature, then $K/Z_p$ also admits a metric of
270: positive scalar curvature if $p=2$ or an odd integer \cite{Rosenberg94}.
271: %************
272: It is likely that all six dimensional Calabi-Yau spaces (whether simply 
273: connected or not) admit metrics of positive scalar curvature
274: \cite{McInnes:2001xp}.
275: 
276: Let $K$ be a compact $n$ dimensional
277: manifold admitting metrics of positive scalar curvature,
278: and consider the space
279: of all smooth Riemannian metrics on $K$. There is the following 
280: simple description of
281: this space.
282: %*******************
283: First note that the space of all metrics is connected: Any metric can
284: be continuously deformed into any other metric. This is most easily seen
285: by characterizing each metric at a point by an orthonormal frame, and noting
286: that one set of linearly independent vectors can certainly be continuously 
287: connected to another. Second,
288: under a conformal rescaling $\tilde g_{mn} = 
289: \psi^{4/(n-2)} g_{mn}$ the scalar curvature transforms as
290: \be\label{rescale}
291: \tilde \R = \psi^{-4/(n-2)} [\R - a \psi^{-1} \nabla^2 \psi ]
292: \ee
293: where $a=4(n-1)/(n-2)$.
294: Now consider the lowest eigenvalue of the conformally
295: invariant Laplacian: 
296: \be
297: -a\nabla^2 \psi + \R\psi = \lambda_0 \psi. 
298: \ee
299: The corresponding eigenfunction is nonzero and can be used as a conformal
300: factor in (\ref{rescale}). The resulting scalar curvature has the same sign
301: as the eigenvalue $\l_0$.
302: Although this
303: construction does not yield constant scalar curvature, the Yamabe conjecture
304: and subsequent proof \cite{Lee87} shows that one can always
305: conformally rescale these metrics to ones with
306: constant scalar curvature.
307: This is very useful since $\l_0$ is just a function on the space of metrics
308: and divides it up into two regions depending on whether the metric can
309: be rescaled to positive scalar curvature or not. 
310: 
311: By continuity, the boundary of the region with $\R>0$ contains metrics
312: with $\R=0$. In addition, there
313: is the possibility of ``islands of $\R=0$" metrics a finite distance away
314: from the positive scalar curvature metrics (The subspace of zero scalar 
315: curvature metrics need not be connected).  
316: However, those metrics must be Ricci flat, $\R_{mn} =0$. To see this, note that
317: if $\R_{mn} \ne 0$ we can perturb the  metric
318: by a multiple of the Ricci tensor,
319: $\delta g_{mn} = b \R_{mn}$. Since $\R_{mn}$ is traceless and divergence
320: free, the change in the scalar curvature under this perturbation is
321: $\delta \R = -b \R_{mn} \R^{mn}$. So by changing the sign of the coefficient
322: $b$, we can construct nearby
323: metrics with either positive or negative scalar curvature.
324: This shows that any metric with zero scalar curvature but nonzero Ricci
325: tensor, must lie on a boundary between positive and negative $\R$ 
326: metrics.\footnote{This is one way to see that the only metrics on $K3$ with
327: zero scalar curvature are the Ricci flat metrics.}
328: 
329: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
330: \begin{figure}[htb]
331:      \centerline{\epsfxsize=8.6cm 
332:        {\epsfysize=6.0cm
333:              \epsffile{fig1.eps}}}
334:       \caption{The space of metrics for $K$ is shown, where $K$ is  e.g. 
335: a simply connected Calabi-Yau 
336: or $G_2$ manifold. The moduli space of Ricci-flat metrics is surrounded by
337: metrics that can be conformally rescaled to have $\R<0$. }
338: \label{1}
339: \end{figure}
340: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
341: 
342: A Ricci flat metric could also lie on the boundary of a region of $\R>0$ 
343: metrics. So a natural question to ask is: Where is the moduli space
344: of Ricci flat metrics in the usual supersymmetric compactification?
345: We now show that they must lie a finite distance away from the 
346: positive scalar curvature metrics (see Fig.~\ref{1}).
347: Supersymmetry guarantees that there are no tachyons in the perturbative
348: spectrum. However all Ricci flat metrics which can be perturbed to
349: both positive and negative scalar curvature must have tachyons.
350: This can be seen as follows.
351:  A general perturbation can be divided into a pure
352: trace, and transverse traceless pieces. A pure trace perturbation produces
353: a first order change in the scalar curvature proportional to $\nabla^2 h$.
354: This integrates to zero so it clearly cannot have definite sign.
355: A transverse traceless perturbation, $h_{mn}$,
356: does not change $\R$ to first order. To study its second order effects,
357: consider the functional
358: \beq
359: S=\int_K \R \sqrt g.
360: \eeq
361: Its first variation is
362: \beq
363: \delta S = \int_K (\R_{mn} - {1\over 2} \R g_{mn})h^{mn} \sqrt g. 
364: \eeq
365: This vanishes for all perturbations since the background is Ricci flat.
366: The second variation is 
367: \beq
368: \delta^2 S = \int_K h^{mn} \Delta_L h_{mn} \sqrt g
369: \eeq
370: where $\Delta_L$ is the Lichnerowicz operator coming from the linearized
371: Einstein equation. It is now clear that $\Delta_L$ must have both positive
372: and negative eigenvalues in order for $\delta^2 S$ to have either sign.
373: But this means that if we start with the product spacetime
374: $M_4\times K$ with metric $ds^2 = \eta_{\mu\nu} dx^\mu dx^\nu
375: + g_{mn}(y) dy^m dy^n $, some metric perturbations of the form
376: $\phi(x) h_{mn}(y)$ will be tachyonic. 
377: Thus the moduli space of supersymmetric, Ricci flat metrics must
378: lie a finite distance\footnote{There probably exist manifolds $K$ that admit 
379: Ricci flat metrics on the boundary of an $\R>0$ region, but they are necessarily
380: not supersymmetric.} away from the metrics with $\R>0$, as shown in Figure 1.
381: 
382: \setcounter{equation}{0}
383: \section{Negative Energy Density in Four Dimensions}
384: 
385: As described in the introduction, the existence of metrics of positive
386: scalar curvature on $K$ leads to effective negative energy density 
387: in the four dimensional theory. In this section, we explain this in
388: more detail and describe some qualitative features of the four dimensional
389: effective potentials which arise.
390: 
391: We want to consider vacuum solutions which asymptotically approach
392: $ M_4 \times K$,
393: where $K$ is, e.g., a simply-connected Calabi-Yau space.
394: We can characterize these solutions in terms of initial data. To minimize
395: the mass, we will consider time symmetric initial data which are
396: spherically symmetric on $\RR$. There are many ways to construct
397: solutions to the initial value  constraint $\R=0$ which have regions of
398: negative energy density on $\RR$. We now describe one approach.
399: 
400: Consider the metric
401: \be \label{ansatz}
402: ds^2 = \left( 1-{2m(r)\over r} \right)^{-1} dr^2 +r^2 d\Omega_2^2
403: +g_{mn} (r,y) dy^{m} dy^{n}  
404: \ee
405: where the indices $m,n$ label the extra compact dimensions.
406: The metric $g_{mn} (r,y)$ denotes a one parameter family of metrics on $K$. 
407: The Ricci scalar of (\ref{ansatz}) is 
408: \beq \label{rscalar}
409: {\cal R} & = & 
410: - \left( 1-\frac{2m(r)}{r} \right)\left[-\frac{1}{4} \partial_{r} g^{mn}
411: \partial_{r} g_{mn} + g''
412: +\frac{2}{r} g' +
413: \frac{1}{4} (g')^2 \right]\nonumber\\
414: & &  
415: + \R_{K}  
416: +\partial_{r} m \left( \frac{4}{r^2} +\frac{g'}{r}  
417: \right)
418: -\frac{m}{r^2} g' 
419: \eeq
420: where $g' \equiv g^{mn} \partial_r g_{mn}$, $g'' = \partial_r g'$, and
421: $\R_{K} $ is the scalar curvature of $g_{mn}(r,y)$ at fixed $r$.
422: 
423: Inside some region $r<R_0$, we choose $g_{mn} (r,y)$ to be independent of
424: $r$, and equal to some metric with $\R_K=2V_0$, a positive constant.
425: In this case, the constraint
426: $\R=0$  reduces to $\partial_r m = -V_0 r^2/2$ which is easily solved
427: for $m(r)$ yielding a region of constant negative energy density. We now
428: pick a radius $R_1>R_0$ and choose any path in the space of metrics which 
429: connects our positive scalar curvature metric $g_{mn}(R_0,y)$ 
430: to a metric on the moduli space, $g_{mn}(R_1,y)$. In general, we cannot 
431: solve $\R=0$ for $m(r)$ because there is nontrivial $y$ dependence. 
432: However, we can
433: find an $m(r)$ so that $\R \ge 0$. We can either view this as nonvacuum
434: initial data for string theory, by adding say a dilaton with $\varphi=0$ and
435: $\dot \varphi^2 = \R$, or we can obtain vacuum initial data by a subsequent
436: conformal rescaling of 
437: the nine dimensional metric (\ref{ansatz}). Let $\widetilde
438: {ds}^2 = e^{4\psi/7} ds^2$, then  the change in the scalar curvature is
439: given by (\ref{rescale}).
440: So if $\psi$ is a solution to the conformally invariant
441: Laplace equation in nine dimensions
442: \be
443: -\nabla^2 \psi + {7\over 32} \R\psi = 0
444: \ee
445: then the rescaled  scalar curvature vanishes. In order for the rescaled metric
446: to be nonsingular and asymptotically flat, we need a solution $\psi$
447: which is nonvanishing and goes to one at infinity. One can show that
448: such solutions always exist when $\R\ge 0$ \cite{Witt}.
449: This conformal rescaling can only decrease the total energy since
450: the ADM mass changes by
451: \be
452: \Delta M \propto -\oint \nabla \psi \propto -\int \R\psi <0. 
453: \ee
454: 
455:  From the four dimensional viewpoint, the metric on $K$ is like an
456:  %*************
457:  (infinite)
458:  collection
459: of scalar fields with potential $-\R_K$. 
460: Qualitatively, this potential has a local minimum at zero when $g_{mn}$
461: is on the moduli space. There is then a finite positive barrier separating
462: this minimum 
463: from a region where the potential is negative. Since the potential is just
464: the scalar curvature, the height of the barrier is roughly $1/L_K^2$ where 
465: $L_K$ is a characteristic size of $K$. Thus large Calabi-Yau spaces have small
466: potential barriers. The width of the potential is harder to estimate
467: since it depends on mathematical details about the space of metrics on $K$
468: which are not yet known.
469: For example, a key open question is: How close does the 
470: moduli space of Ricci flat metrics come to the region of positive scalar
471: curvature metrics? The positive energy theorems we discuss later can be
472: used to give some information about this distance.
473: 
474: Once one reaches a metric of constant scalar curvature $\R_K =2V_0>0$, one can
475: always rescale the metric by a conformal factor which is
476: constant on $K$, to increase the
477: curvature and make the effective three dimensional energy density more
478: negative. We can easily compute the effective potential for this mode.
479: Let us start with a product metric on $\RR \times K$, $ds^2 = ds_3^2 +
480: ds^2_K$. Let $\phi$ be a function depending only on $\RR$. The scalar
481: curvature of the metric
482: \be
483: ds^2 = e^{-n\phi} ds^2_3 + e^{2\phi} ds_K^2
484: \ee
485: is
486: \be
487: \R = e^{n\phi}[\R_3 +2V_0 e^{-(2+n)\phi} - {n(n+2)\over 2} (\nabla \phi)^2]
488: \ee
489: where $n$ is the dimension of $K$.
490: The second term on the right is just the scalar curvature of the rescaled
491: metric on $K$.
492: The vacuum constraint is $\R=0$, and in $3+1$ dimensions, the energy density
493: is $\rho =\R_3/2$. So we obtain
494: \be 
495: \rho = {n(n+2)\over 4} (\nabla \phi)^2 - V_0 e^{-(2+n)\phi}. 
496: \ee
497: Rescaling $\phi$ to have a standard kinetic term we get
498: \be
499: \rho = {1\over 2}(\nabla \tilde \phi)^2 - V_0 e^{-\alpha \tilde \phi} 
500: \ee
501: where 
502: \be\label{defalpha}
503: \alpha^2 = {2(n+2)\over n}.
504: \ee 
505: So the potential is not only negative, but
506: falls off exponentially fast. Notice that for more than one extra dimension,
507: $2<\alpha^2<6$.
508: 
509: There are other ways to achieve large $\R_K$ (and hence large negative
510: energy density). We mentioned in the last section
511: that one can always conformally rescale
512: a metric on $K$ to one with constant scalar curvature whose sign 
513: depends on the lowest eigenvalue, $\lambda_0$, of the conformal Laplacian.
514: If $\lambda_0 \le 0$, the rescaled metric is unique for fixed  volume.
515: In other words, the only remaining freedom is the
516: constant rescalings discussed above. However, if $\lambda_0 >0$, the
517: conformally related metric with constant positive curvature is generically
518: not unique,
519: even when the volume is fixed.
520: There is a maximum value of the scalar curvature in each conformal
521: equivalence class, but it can be arbitrarily big in nearby conformal classes.
522: More precisely, given a conformal class with $\lambda_0 >0$, for any
523: $\varepsilon>0$ and any large number $N$ there is another conformal class
524: within $\varepsilon$  of the first (in a $C^0$ topology) with a metric
525: of unit volume and constant
526: scalar curvature larger than $N$ \cite{Pollack}. This surprising result
527: shows that the maximum scalar curvature in each conformal class is not
528: a continuous function\footnote{This depends on the topology one puts on the
529: space of conformal classes. In a stronger topology, this function will probably
530: be continuous \cite{Pollack2}.}.
531: 
532: There is clearly a lot of freedom in the construction of configurations
533: with negative energy density. One can choose any path in the space of
534: metrics from the moduli space to the region with $\R_K>0$,
535: and then choose any path in the region of positive scalar curvature metrics
536: to make $\R_K$ large. One approach is to divide the motion up into
537: nonconformal motion and conformal rescalings. For the nonconformal motion,
538: there are two natural choices:
539:  One is to choose a path that only
540: includes metrics of constant scalar curvature and fixed volume; the other
541: is to keep a local volume element fixed. This second option greatly simplifies
542: (\ref{rscalar}) since $g'=0$, but since the scalar curvature does not
543: remain constant, one needs to rescale the final metric by
544: a position dependent conformal factor to obtain 
545: constant $\R_K>0$.
546: 
547: \setcounter{equation}{0}
548: \section{Scalar Field Model}
549: 
550: In the next section we will review the theorems which show that
551: despite the unbounded negative energy density,
552: the total energy must remain positive.
553: To gain intuition into how this is possible,  we now
554: consider a simple model of four dimensional gravity coupled to a
555: single scalar field with potential $V(\phi)$.
556: To model the situation
557: we have for the compactifications, we
558: consider  potentials $V(\phi)$ which are negative in a region of $\phi$
559: which is separated by a positive barrier from  a local minimum at
560: $\phi_f >0$ where the potential vanishes.
561: We consider
562: configurations where $\phi \rightarrow \phi_f$ asymptotically,
563: and ask whether the total energy can be negative. In the absence of 
564: gravity, the answer is clearly yes: Inside a large ball of radius $R_0$,
565: one can 
566: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
567: \begin{figure}[htb]
568:      \centerline{\epsfxsize=9.6cm 
569:        {\epsfysize=7.5cm
570:              \epsffile{fig2.eps}}}
571:       \caption{A potential with a negative minimum at $\phi=0$ that   
572: is separated by a positive barrier from a local minimum at $\phi_f$.}
573: \label{2}
574: \end{figure}
575: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
576: keep $\phi$ constant at a value where the potential is negative, and
577: then have $\phi$ approach $\phi_f$ in a transition region. Since the
578: negative energy in the ball grows like the volume and the positive energy
579: in the transition region only grows like the area, the total energy
580: can clearly be negative. However, as first shown by Coleman and De Luccia
581: \cite{Coleman:1980aw},
582: gravity can modify this picture.
583: 
584: Since we want to minimize the energy, we set all the time derivatives
585: to zero. 
586: For time symmetric initial data the constraint equations reduce to
587: \be
588: \ ^{(3)}{\cal R} = 2 \rho
589: \ee
590: where
591: \be
592: \rho = \frac{1}{2} g^{ij}\phi_{,i}\phi_{,j} +V(\phi). 
593: \ee
594: Since spatial gradients raise the energy,
595: we consider a spherically symmetric configuration with metric
596: \be
597: ds^2 = \left(1-{2m(r)\over r}\right)^{-1} dr^2 + r^2 d\Omega^2_2. 
598: \ee
599: The constraint then
600: yields the following equation for the ``mass" $m$ as a function of the radius,
601: \be \label{mscalar}
602: m' +\frac{1}{2} mr\phi'^2 = \frac{1}{2} r^2 \left[V(\phi)+\frac{1}{2}\phi'^2
603: \right]
604: \ee
605: where prime denotes the ordinary derivative with respect to $r$.
606: The ADM mass is simply the value of $m(r)$ at infinity.
607: 
608: We can choose $\phi(r)$ arbitrarily and solve (\ref{mscalar}) for $m(r)$. 
609: The general solution is
610: \be\label{gensoln} 
611: 2m(r) =  e^{-\int_{0}^{r} \hat r\phi'^2/2\, d\hat r} 
612: \int_{0}^{r} e^{\int_{0}^{\tilde r} \hat r\phi'^2/2\, d\hat r} \left[V(\phi) +
613: {1\over 2}\phi'^2 \right] 
614: \tilde r^2 d\tilde r.  
615: \ee
616: %Notice the exponential suppression in front of the integral and the
617: %exponential enhancement inside. We will see that these factors 
618: %have a big effect on the mass.
619: Consider now a potential with a negative 
620: minimum at $\phi =0$ that is separated by a positive barrier
621: from a local minimum at $\phi_f $, 
622: where the potential vanishes.
623: We call the height of the positive barrier $V_1$, and the depth of the
624: negative minimum $V(0)=-V_0$ 
625: (see Fig.~\ref{2}). 
626: We pick two
627: radii $R_1>R_0$ and  set
628: $\phi=0$ for $r<R_0$, $\phi= \phi_f$ for $r>R_1$,
629: and let $\phi$ be any smooth increasing function for $R_0<r<R_1$.
630: The solution for $m(r)$ for this configuration is 
631: \be 
632: 2m(r) = e^{-\int_{R_0}^{r} \hat r\phi'^2/2} \left[-\frac{1}{3} V_0 R_0^3
633: +\int_{R_0}^{r} \left (V(\phi) +
634: {1\over 2}\phi'^2 \right ) e^{\int_{R_0}^{\tilde r} \hat r\phi'^2/2} 
635: \tilde r^2 d\tilde r \right]. 
636: \ee
637: One sees that the positive energy in the barrier is enhanced relative
638: to the negative energy by an exponential factor.
639: We shall see that the ADM mass can nevertheless be negative provided
640: that $V_0/V_1$ is sufficiently large. To identify the approximate criteria
641: for this to be the case, we minimize the exponential factor
642: by taking 
643: \be
644: \phi (r) = \mu \ln (r/R_0)
645: \ee
646: in the transition region. This fixes the relation between $R_0$ and $R_1$
647: to be
648: $R_1=R_0 e^{\phi_{f}/\mu}$. The ADM mass $M=m(\infty)=m(R_1)$ is given by
649: \beq
650: 2M &= & e^{- \mu \phi_{f}/2} \left[ -\frac{1}{3}V_0 R_0^3
651: +\frac{\mu^2 R_0}{( \mu^2 +2)} \left(e^{\phi_{f}(\mu^2 +2)/2\mu}-1 \right)
652:  \right.\nonumber\\
653: & & +\left. \frac{R_0^3}{\mu} \int_{0}^{\phi_{f}} e^{x(\mu^2+6)/2\mu}V(x)dx 
654: \right]. 
655: \eeq
656: Notice that the last term, coming from the transition region,
657: scales like $R_0^3$, just like the first term. 
658: If the slope $\mu$  and the width of the potential barrier $\phi_f$ are
659: $ \sim {\cal O}(1)$, then the integral is $ \sim V_1$.
660: For $R_0 \gg 1$, the condition for the ADM mass $M$ to be negative is 
661: then simply $V_{0} >V_1$.
662: For an arbitrary slope $\mu$ and width  $\phi_f$,
663: and assuming $V(\phi) =V_1$ for 
664: $0<\phi<\phi_{f}$, we have $M<0$ if (again for $R_0\gg 1$)
665: \be
666: \frac{V_0}{V_1} > \frac{6}{(\mu^2 +6)}
667: \left(e^{\phi_{f}(\mu^2+6)/2\mu}-1 \right). 
668: \ee
669: The right hand side is minimized for $\mu$ of order one.
670: 
671: Intuitively, one reason that it is harder to get negative total energy
672: in theories with gravity is that a constant negative energy
673: density produces a hyperbolic space with constant negative curvature. In this
674: space, the volume inside a ball of large radius is equal to 
675: the volume of a thin shell at this radius with a width equal to the scale of
676: curvature. 
677: 
678: As we saw in the last section, the potentials which arise
679: from Calabi-Yau compactifications are unbounded from below.
680: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
681: \begin{figure}[htb]
682:      \centerline{\epsfxsize=9.6cm 
683:        {\epsfysize=8.0cm
684:              \epsffile{fig3.eps}}}
685:       \caption{A potential which is unbounded from below for 
686:  negative values of $\phi$.}
687: \label{3}
688: \end{figure}
689: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
690: Now suppose we modify the potential for negative $\phi$ so that it
691: falls off exponentially $V = - V_0 e^{-\alpha\phi}$~(see Fig.~\ref{3}). 
692: Since this potential is unbounded from below, one might expect
693: that one can always construct negative energy solutions. But this is not
694: the case. We now show that there is a critical value,  $\alpha^2 =6$, such
695: that if $\alpha^2<6$ the energy can stay positive by an order one 
696: barrier. 
697: %This can be viewed as an extreme version of Coleman De Luccia's
698: %observation that gravity can stabilize a false vacuum.
699: 
700: As we saw above, the key point is how low can we make the energy density
701: inside a ball of radius $R_0$, if we require $\phi(R_0)=0$. If we keep
702: $\phi=0$ everywhere inside the ball, the energy density would be $-V_0$.
703: If we include negative values of $\phi$, we can access the growing negative
704: potential, but we have to pay the price of the derivative contributions
705: to the energy and the exponential enhancement of the barrier.
706: Clearly, to make the
707: energy density as negative as possible, we want $\phi$ to change
708: slowly. This minimizes the positive $\phi'^2$ contribution to the
709: energy and increases the region over which the negative potential
710: contributes. However, the most important effect is the exponential
711: factor. To minimize the exponential, we take
712: $\phi= \mu\ln (r/R_0)$. This configuration is singular at the origin, but
713: we will smooth it out by setting $\phi$ equal to a constant $ -\phi_0 <0$
714: for $r<r_0$. Thus $\phi_0 = \mu \ln (R_0/r_0)$. Since $\phi$ is constant,
715: the exponential vanishes for $r<r_0$ and for $r_0 < r< R_0$ we have
716: \be
717:  \int_{r_0}^r r\phi'^2 = \mu^2 \ln (r/r_0). 
718: \ee
719: Using the general solution (\ref{gensoln}) we get
720: \beq
721: 2m(R_0)& =& -{ r_0^3V_0\over 3} \({r_0\over R_0}\)^{{\mu^2\over 2} -\alpha\mu}
722:    -{2V_0\over\mu^2 -2\alpha\mu +6}
723:    \left [ R_0^3 - r_0^3 \({r_0\over R_0}\)^{{\mu^2\over 2} -\alpha\mu}\right]
724:    \nonumber\\
725:  &  &+{\mu^2\over \mu^2 +2}
726:    \left [ R_0 - r_0 \({r_0\over R_0}\)^{\mu^2/2}\right]. 
727: \eeq
728: The first term is the contribution to the energy from the inner region
729: $r<r_0$, the second term is the contribution from the potential in
730: the region $r_0 <r<R_0$, and the last term is the contribution from the
731: derivative terms. To make the energy density inside $R_0$ as negative as
732: possible, we should choose $r_0 \ll R_0$. Then 
733: \be\label{negdensity}
734: {m(R_0)\over  R_0^3} =- {V_0\over \mu^2 - 2\alpha\mu +6}. 
735: \ee
736: For $\alpha^2<6$, this is maximized at $\mu=\alpha$, when the right hand
737: side becomes
738: $-V_0/(6-\alpha^2)$. If $\alpha^2 \ge 6$, the energy
739: density inside $R_0$  can be made arbitrarily negative, by choosing $\mu$
740: so that the denominator in (\ref{negdensity}) is as small as one wants. In 
741: this case, no finite positive barrier can prevent the total ADM energy from 
742: being negative. Recall that the values of $\alpha$ that arise from 
743: compactification (\ref{defalpha}) all satisfy $\alpha^2 <6$.
744: 
745: One can show that $\phi(r) = \alpha \ln (r/R_0)$ is indeed an extremal path
746: for the potential
747: energy, given an exponential potential $V = -V_0e^{-\alpha\phi}$.
748: If one changes this path by a small perturbation $\varepsilon(r)$ with 
749: $\varepsilon(R_0)=0$,
750: then the $R_0^3$ contribution to the energy does not change to first order
751: in $\varepsilon$.
752: 
753: 
754: 
755: 
756: \setcounter{equation}{0}
757: \section{Positive energy theorems}
758: 
759: Despite the fact that the energy density can be arbitrarily negative,
760: the total energy must remain positive. This follows from 
761: a simple extension of Witten's proof \cite{Witten:mf}
762: of the positive energy theorem\footnote{The proof by Schoen and Yau
763: \cite{SY} does not extend to ten or eleven dimensions.}. We first review the
764: original four dimensional version, and then discuss its extension 
765: to higher dimensions. Consider a 
766: four dimensional  asymptotically flat
767: spacetime with matter satisfying the dominant energy
768: condition: $T_{\mu\nu} t^\mu \tilde t^\nu \ge 0$, where $t^\mu$ and
769: $\tilde t^\nu$ are any two future directed timelike vectors. Let $\Sigma$
770: be a nonsingular, asymptotically flat spacelike surface and let $D_i$
771: be the projection of the spacetime covariant derivative into $\Sigma$.
772: Let $\e$ be a solution of the spatial Dirac equation $\gamma^i D_i \e=0$
773: which approaches a constant spinor $\e_0$ at infinity. Then one can
774: show that 
775: \be\label{pet}
776: \oint_\infty \e^\dagger D_i \e\ dS^i = \int_\Sigma [(D_i \e)^\dagger (D^i \e)
777: + {1\over 2}T_{\mu\nu} n^\mu (\e^\dagger \gamma^\nu \e)] d\Sigma
778: \ee
779: where $n^\mu$ is the unit normal to $\Sigma$. The right
780: hand side is nonnegative, and the surface integral at infinity is proportional
781: to $P_\mu t_0^\mu$ where $P_\mu$ is the total
782: ADM four momentum and $t_0^\mu =\e^\dagger_0 \gamma^\mu \e_0$ is a timelike
783: (or null) vector. This proves that the total ADM energy cannot be negative.
784: If the total energy vanishes, $T_{\mu\nu}=0$, $\e$ is covariantly constant,
785: and the spacetime must be Minkowski space.
786: The only subtlety in the proof is showing that asymptotically 
787: constant solutions to $\gamma^i D_i \e=0$ exist. A plausible argument
788: was given by Witten and  a rigorous proof was given in \cite{Parker:uy}.
789: 
790: This argument easily extends to 
791: higher dimensional spacetimes which admit spinors and
792: have a covariantly constant spinor at infinity. The calculation
793: which leads from  $\gamma^i D_i \e=0$ to (\ref{pet}) continues to hold
794: in higher dimensions. Since supersymmetric vacuum compactifications $K$ 
795: always admit covariantly constant spinors,  solutions
796: which asymptotically approach $M_4\times K$ cannot have negative total energy.
797: Furthermore, the only solution with zero energy is $M_4\times K$ itself.
798: Once again, the only
799: subtlety is the existence of appropriate 
800: solutions to the spatial Dirac equation.
801: For Calabi-Yau compactifications, this has recently been shown rigorously
802: by Dai \cite{Dai}. It seems surprising that this proof of positive total energy
803: treats K3 and Calabi-Yau compactifications identically, even though we
804: have seen that the latter have
805: four dimensional regions of negative energy density and the former
806: do not. On a time symmetric surface $\Sigma =\RR \times K$ in a vacuum
807: solution, the proof only uses the
808: fact that the scalar curvature of $\Sigma$ vanishes, and does not care
809: whether there are positive scalar curvature metrics on $K$  or not.
810: In fact, Witten's proof of positive ADM energy was given before it was even
811: known that such metrics of positive scalar curvature exist.
812: 
813: %*********************
814: %One can also show that the ADM energy remains positive
815: From a purely 
816: four dimensional viewpoint,
817: Calabi-Yau compactifications
818: include fields with
819: potentials that are unbounded from below. These certainly do not satisfy
820: the dominant energy condition, so the usual form of the
821: %*************
822: (four dimensional)
823: positive energy
824: theorem does not apply. It is well known that for
825: asymptotically anti de Sitter spacetimes, one can have positive energy even
826: with potentials which are unbounded from below 
827: \cite{Breitenlohner:jf,Gibbons:aq}.  
828: It is much less known that 
829: this is also possible in asymptotically flat spacetimes.
830: However this was shown almost twenty years ago by 
831: Boucher \cite{Boucher:1984yx} (see also \cite{Townsend:iu}).
832: The idea is to modify the
833: covariant derivative in Witten's spatial
834: Dirac equation to include an arbitrary function of the fields.
835: In the simplest case of a single scalar field, one takes
836: $\hat D_i = D_i + i W(\phi) \gamma_i $. It then turns out that
837: one can prove positive energy provided the potential for $\phi$
838: takes the form 
839: \be\label{superpot}
840: V = 2 W'^2 - 3W^2. 
841: \ee
842: This is similar to the form of the scalar potential in $N=1$ supergravity,
843: in terms of the superpotential $W$, but one can do this even in 
844: nonsupersymmetric theories. 
845: This potential $V$ can be unbounded from below, but if it has a local
846: minimum at $V=0$, then there will be asymptotically flat solutions and
847: the positive energy theorem will hold.
848: 
849: We do not know the exact form of the potentials which arise in Calabi-Yau
850: compactifications far off the moduli space.   
851: If we start with Type II string theory, we obtain
852: four dimensional $N=2$ supergravity coupled to matter, and
853: previous analysis has mostly focused
854: on the massless moduli fields. The positive energy theorem can be
855: used to say something about the shape of the potential away from the
856: moduli space. Consider
857:  the simple conformal mode that we discussed in section 3 with
858: exponential potential. Recall that
859: the potential was $V= - V_0 e^{-\alpha \phi}$, with $\alpha^2 <6$.
860: This applies inside the region of positive scalar curvature. Moving from
861: this region to the moduli space corresponds to crossing a potential
862: barrier of height $\sim 1/L_K^2$, where $L_K$ is the size of the
863: Calabi-Yau
864: space.
865: A positive energy theorem is possible only if the width of the potential
866: is large enough. If $V_0 \sim 1/L_K^2$, the width must be at least
867: of order one.
868: 
869: It was suggested in \cite{DeWolfe:1999cp} that one could take any $V(\phi)$ and solve
870: (\ref{superpot}) for $W(\phi)$, but clearly global solutions only exist for a
871: restricted class of $V(\phi)$. We saw in section 4 that 
872: the exponential potential
873: $V= - V_0 e^{-\alpha \phi}$ with $\alpha^2 \ge 6$ always has negative
874: energy solutions and hence a solution for $W$ cannot exist. In contrast,
875:  for $\alpha^2 < 6$, one can obtain $V$ from 
876:  $W= W_0 e^{-\alpha\phi/2}$ with $W_0^2 = 2 V_0/(6-\alpha^2)$.
877: Note that 
878: the critical value of the exponent which arose in constructing negative
879: energy solutions is precisely the one which allows  a global
880: solution for $W$.
881: 
882: \section{Physical consequences of negative energy density}
883: 
884: 
885: Even though the total energy stays positive, the existence of
886: unbounded negative energy density may have serious consequences.
887: In this section we discuss some of these.
888: 
889: \subsection{Cosmic Censorship}
890: 
891: The existence of negative energy density may lead to violations of
892: cosmic censorship. That is, one might have nonsingular initial data,
893: which evolves to singularities that are not hidden inside event
894: horizons. As an example, consider a theory with a 
895: single scalar field and negative
896: exponential potential  $V(\phi) = - e^{\alpha\phi}$ for $\phi>0$
897: and a local minimum at $\phi_f<0$ where $V=0$. (This is similar to the
898: example in section 4 except that the sign of $\phi$ has been changed to
899: remove some unnecessary minus signs.)
900: Consider time symmetric initial data consisting
901: of a constant scalar field $\phi= \phi_0 >0$ for $r<R_0$ and then
902: $\phi$ changes continuously to the local minimum of the potential at $V=0$
903: at larger radius. Inside the sphere $r=R_0$, the energy density is
904: a negative constant, so the three geometry has constant negative curvature.
905: Thus the evolution inside the domain of dependence of the ball of radius
906: $R_0$ looks like a $k=-1$ Robertson-Walker universe. 
907: (It is NOT anti de Sitter space, since the scalar field evolves.)
908: Under evolution, the scalar field rolls down the potential and the
909: spacetime becomes singular in finite time.
910: Since this singularity is 
911: not the result of the usual gravitational collapse, one may not form trapped
912: surfaces or an event horizon. This would lead to a violation of
913: cosmic censorship.
914: 
915: To examine this possibility further, let us first estimate the time to the
916: singularity. Since the metric takes  the
917: standard Robertson-Walker form, 
918: $ds^2 = -dt^2 + a^2(t) d\sigma^2$ where $d\sigma^2$ is the metric on
919: the unit hyperboloid,   the field equations are 
920: \be\label{fieldeq}
921: {\ddot a\over a}= {1\over 3}[V(\phi) - \dot\phi^2]
922: \ee
923: \be
924: \ddot{\phi} +{3 \dot a\over a} \dot\phi - \alpha e^{\alpha\phi} =0. 
925: \ee
926: We start with $\phi=\phi_0$, and set $\dot\phi=0, \dot a=0$ since
927: we want time symmetric initial data.  The right hand side
928: of (\ref{fieldeq}) is negative, so $a$ decreases under evolution and the 
929: singularity occurs when $a=0$. Since $\phi$ increases 
930: as it rolls down the potential, the right hand side of (\ref{fieldeq})
931: is always less than its initial value. Thus we can get an upper bound
932: on the time to the singularity by solving the simple equation
933: $\ddot a + a e^{\alpha \phi_0}/3  =0$.
934: This yields 
935: \be 
936: T < \frac{\sqrt 3\pi}{2} e^{-\alpha\phi_0/2}. 
937: \ee
938: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
939: \begin{figure}[htb]
940:      \centerline{\epsfxsize=9.6cm 
941:        {\epsfysize=5.5cm
942:              \epsffile{fig4.eps}}}
943:       \caption{If an event horizon encloses the singularity, it must have
944:       an initial size of order $R_0$.}
945: \label{4}
946: \end{figure}
947: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
948: 
949: If we choose $R_0 \gg T$, then the singularity will form very quickly
950: compared to the size of the region. If the singularity lies inside
951: a black hole, then we can trace the null geodesic generators of the event 
952:  horizon back to the initial surface where it will form a sphere of radius
953: approximately $R_0$ (see Fig.~\ref{4}).
954: Now suppose the total mass is smaller than a Schwarzschild
955: black hole of size $R_0$. 
956: The area theorem for black holes only requires the weak energy condition
957: and hence still holds even in theories with $V(\phi)<0$.
958: Since the area of the event horizon
959: cannot decrease during evolution and the mass cannot increase,
960: when the spacetime settles down there is  not enough mass to support
961: a black hole large enough to enclose the singularity.
962:  Inside the domain of
963: dependence of the initial ball of radius $R_0$,
964: the singularity will be spacelike like a big crunch. The singularity
965: is likely to extend outside the domain
966: of dependence.
967: There are two
968: possibilities. Either
969: the singularity ends, and the endpoint is a naked singularity, or the 
970: singularity continues all the way out to null infinity cutting off all
971: spacetime. This is a disaster much worse than naked singularities.
972: 
973: Support for this line of thinking comes from the fact that cosmic 
974: censorship is violated in certain four dimensional field theories of 
975: this type \cite{Hertog03}.
976: We do not yet know if this is realized in Calabi-Yau compactifications, but
977: it cannot be ruled out.
978: %*********** 
979: In the scalar field example, one can show from the field equations that
980: $\phi$ diverges at the singularity. In the Calabi-Yau context, this would
981: correspond
982: to $K$ shrinking to zero size. If one tries to do this keeping the
983: metric on $K$ Ricci flat, one is likely to form black holes 
984: \cite{Geddes:2001ht}. Since our data are time symmetric, there will be
985: singularities in the past as well as the future. If naked singularities
986: form in this case, it would be interesting to see if they also
987: form in situations without singularities in the past.
988: 
989: Violations of cosmic censorship could be viewed as a desirable feature
990: of Calabi-Yau compactifications. If singularities can be visible, one
991: has the possibility of directly observing effects of quantum gravity.
992: Of course, the key question is how frequently do naked singularities form.
993: If they occur too often in certain compactifications, 
994: then those would be observationally 
995: ruled out\footnote{It is not difficult to see that cosmic censorship can be 
996: violated in nonsupersymmetric compactifications, $M_4\times K$, 
997: if $K$ is a manifold that admits a Ricci flat metric on or close to the
998: boundary of an $\R_K>0$ region. Of course, such theories are physically 
999: unacceptable anyway since they do not admit a positive energy theorem.}.
1000: 
1001: \subsection{New thermal instability}
1002: 
1003: Since there is only a finite barrier separating the 
1004: supersymmetric vacuum $M_4\times K$
1005: from the region of unbounded negative energy density, one can
1006: ask whether ordinary finite energy processes can cause an instability.
1007: It is unlikely that a two particle collision at high energy can trigger
1008: this instability. This is because the classical process of the field going
1009: over the barrier is described quantum mechanically by a coherent state
1010: which involves many quanta. However, it is possible to see this
1011: instability in a thermal state at nonzero temperature.
1012: 
1013: At nonzero temperature, there is always an instability for nucleating a 
1014: black hole \cite{Gross:cv}. There is a gravitational instanton describing this
1015: process semiclassically. The instanton is the euclidean Schwarzschild
1016: black hole with the prescribed temperature, and the nucleation rate is given
1017: by the action of this instanton.
1018: This yields $\Gamma \sim e^{-M^2} = e^{-1/T^2}$, so the nucleation
1019: of black holes is highly suppressed at temperatures much below the Planck 
1020: scale ($T_{pl} \sim 10^{32} K$).
1021: 
1022: The existence of negative energy density in Calabi-Yau compactifications
1023: means there is a new instability corresponding to thermal fluctuations causing
1024: the field to jump to the top of the barrier in some region of space 
1025: and rolling over\footnote{If the width of the potential is small enough, this
1026: instability might manifest itself in terms of quantum tunneling through
1027: the barrier.}. The height of the barrier is
1028: roughly $1/L_K^2$ where $L_K$ is the size of the Calabi-Yau space.
1029: If only a small region has a thermal excitation of
1030: this amount, it just disperses. But if a region of size $L_K$ undergoes
1031: such a thermal excitation it will become unstable. This is 
1032: because this region will now act like a horizon sized patch of de Sitter.
1033: Typically, the region briefly expands before the field rolls down the 
1034: potential. If the field rolls down on the other side,
1035: the region will collapse down to a singularity
1036: \cite{Banks:1984cw,Felder:2002jk}. 
1037: The energy in the original thermal excitation is 
1038: roughly $E \sim L_K$, so the decay rate is 
1039: \be
1040: \Gamma \sim e^{-L_K/T}.
1041: \ee 
1042: For low temperatures $T<1/L_K$, this is larger than the rate of black hole
1043: production.  One could try to make this estimate more
1044: precise by looking for a new thermal instanton solution. 
1045: In four dimensional gravity coupled to a scalar field with a potential
1046: as in figure 3, the decay of the false vacuum at finite temperature
1047: would be described by an $O(3)$-symmetric gravitational instanton
1048: that is periodic in Euclidean time. In our case this would be a
1049: ten dimensional euclidean vacuum solution which asymptotically approaches
1050: $\RR \times S^1 \times K$. Whether or not this new thermal instability
1051: is significant in the early universe 
1052: depends on the size of the internal space. 
1053: 
1054: 
1055: \section{Discussion}
1056: We have reexamined the four dimensional theories resulting from supersymmetric
1057: compactifications of string theory (or just supergravity). Most previous
1058: discussions have focused on the moduli space of vacua
1059: and the associated massless fields. We have shown that if one moves off the
1060: moduli space, in many cases the four dimensional effective potentials are
1061: unbounded below. This applies to all simply connected Calabi-Yau and $G_2$
1062: manifolds, as well as some nonsimply connected ones.
1063: 
1064: Naively, this suggests that there should be solutions with negative
1065: total ADM energy, but this is not the case. One can view the fact
1066: that the energy of all nonsingular solutions  remains positive as
1067: a generalization of the phenomenon that gravity
1068: can stabilize a false vacuum \cite{Coleman:1980aw}.
1069: 
1070: We have begun an investigation of the physical effects of the negative
1071: energy density, and have discussed possible violations of cosmic censorship
1072: and new thermal instabilities. In Calabi-Yau compactifications, there
1073: is another type of possible violation of cosmic censorship 
1074: %*******************
1075: %which has been discussed.
1076: In the moduli space of Ricci flat metrics, one can deform
1077: a smooth metric into a singular one, e.g., at a conifold point. Since there
1078: are solutions of the form $M_4\times K$ for every point on the moduli space,
1079: one may be able to  change the metric on $K$ slowly enough
1080: so that no event horizon forms in the four noncompact directions even though
1081: a singularity develops on $K$. 
1082: The possible violation of cosmic censorship discussed in the previous section 
1083: is quite
1084: different. In our case, a singularity is forming in the four noncompact
1085: dimensions, not just on $K$. Furthermore, while it has been shown that
1086: the conifold singularity is harmless in string theory, the singularities
1087: we have been discussing may not be. However, it should be kept in mind that
1088: in neither case have violations of cosmic censorship been firmly established
1089: for Calabi-Yau compactifications.
1090: 
1091: 
1092: There may be other physical consequences of the negative energy density.
1093: If these consequences are too severe, they may show that all these
1094: compactifications are unphysical. A longstanding problem in string
1095: theory is the large number of apparently consistent vacua in the theory.
1096: Conceivably, the negative energy density could cut down this number dramatically.
1097: 
1098: Although we have focused on asymptotically flat solutions, the existence
1099: of potentials which are unbounded from below is likely to be important
1100: in other contexts as well\footnote{These applications were suggested
1101: by E. Silverstein and L. Susskind.}. 
1102: For example, the recent construction of de Sitter
1103: space as a solution to string theory  \cite{Kachru:2003aw}
1104: involves a Calabi-Yau internal space. Since
1105: space is now compact, there is no positive energy theorem. The negative energy
1106: density we have found
1107: provides a new type of instability of this de Sitter solution.
1108: As another example,
1109: the existence of negative energy regions
1110: surrounded by positive energy in such a way that the total energy remains
1111: positive is reminiscent of quantum field theory \cite{Ford:1999qv}.
1112: In the context of AdS/CFT
1113: one might wonder whether negative energy density in the (quantum) CFT is 
1114: related to negative energy density in the (classical)  AdS supergravity.
1115: 
1116: Just because $K3$ does not have metrics of positive scalar curvature does
1117: not guarantee that it is free of potential instabilities. 
1118: S-duality relates Type II string theory on certain Calabi-Yau manifolds
1119: to heterotic string theory on $K3\times T^2$ 
1120: \cite{Kachru:1995wm,Ferrara:1995yx}. In some of these cases, it is known
1121: that the Calabi-Yau space
1122: admits a metric with $\R_K>0$. The fact that the Type II
1123: string has regions of negative energy density means that the same thing must
1124: be true for the heterotic string on $K3\times T^2$. Of course the negative
1125: energy density will not arise in the same way, and  may be a
1126: strong coupling effect.
1127: 
1128: Another reason for questioning the stability of $K3$ (as well as Calabi-Yau)
1129: compactifications is some recent work of Penrose \cite{Penrose02}. He
1130: argues that all curved compactifications should  have
1131: the property that generic finite perturbations will produce  curvature
1132: singularities. This is worthy of further investigation.
1133: 
1134: \bigskip
1135: 
1136: \centerline{{\bf Acknowledgments}}
1137: It is a pleasure to thank M. Cvetic, X. Dai, J. Isenberg, S. Rey,
1138: R. Roiban, J. Polchinski, I. Singer, R. Schoen, K. Schleich, K. Skenderis,
1139: M. Srednicki,
1140: J. Walcher, N. Warner, G. Wei, D. Witt, H. Verlinde, and S.T. Yau for 
1141: discussions. We also thank R. Penrose for stimulating our interest in
1142: the possible instability of compact extra dimensions.
1143: This work was supported in part by NSF grant PHY-0070895 and 
1144: a Yukawa fellowship. 
1145: 
1146: 
1147: 
1148: \begin{thebibliography}{99}
1149: 
1150: %\cite{Witten:gj}
1151: \bibitem{Witten:gj}
1152: E.~Witten,
1153: ``Instability Of The Kaluza-Klein Vacuum,''
1154: Nucl.\ Phys.\ B {\bf 195} (1982) 481.
1155: %%CITATION = NUPHA,B195,481;%%
1156: 
1157: %\cite{Brill:di}
1158: \bibitem{Brill:di}
1159: D.~Brill and H.~Pfister,
1160: ``States Of Negative Total Energy In Kaluza-Klein Theory,''
1161: Phys.\ Lett.\ B {\bf 228} (1989) 359;
1162: %%CITATION = PHLTA,B228,359;%%
1163: D.~Brill and G.~T.~Horowitz,
1164: ``Negative Energy In String Theory,''
1165: Phys.\ Lett.\ B {\bf 262} (1991) 437.
1166: %%CITATION = PHLTA,B262,437;%%
1167: 
1168: %\cite{Witten:mf}
1169: \bibitem{Witten:mf}
1170: E.~Witten,
1171: ``A Simple Proof Of The Positive Energy Theorem,''
1172: Commun.\ Math.\ Phys.\  {\bf 80} (1981) 381.
1173: %%CITATION = CMPHA,80,381;%%
1174: 
1175: %\cite{Boucher:1984yx}
1176: \bibitem{Boucher:1984yx}
1177: W.~Boucher,
1178: ``Positive Energy Without Supersymmetry,''
1179: Nucl.\ Phys.\ B {\bf 242} (1984) 282.
1180: %%CITATION = NUPHA,B242,282;%%
1181: 
1182: \bibitem{Dai}
1183: X. Dai, ``A Positive Mass Theorem for Spaces with SUSY Compactifications",
1184: to appear.
1185: 
1186: %\cite{Coleman:1980aw}
1187: \bibitem{Coleman:1980aw}
1188: S.~R.~Coleman and F.~De Luccia,
1189: ``Gravitational Effects On And Of Vacuum Decay,''
1190: Phys.\ Rev.\ D {\bf 21} (1980) 3305.
1191: %%CITATION = PHRVA,D21,3305;%%
1192: 
1193: %\cite{Weinberg:id}
1194: \bibitem{Weinberg:id}
1195: S.~Weinberg,
1196: ``Does Gravitation Resolve The Ambiguity Among Supersymmetry Vacua?,''
1197: Phys.\ Rev.\ Lett.\  {\bf 48} (1982) 1776.
1198: %%CITATION = PRLTA,48,1776;%%
1199: 
1200: %\cite{Cvetic:1992st}
1201: \bibitem{Cvetic:1992st}
1202: M.~Cvetic, S.~Griffies and S.~J.~Rey,
1203: ``Nonperturbative stability of supergravity and superstring vacua,''
1204: Nucl.\ Phys.\ B {\bf 389} (1993) 3
1205: [arXiv:hep-th/9206004].
1206: %%CITATION = HEP-TH 9206004;%%
1207:  
1208: %\cite{Breitenlohner:jf}
1209: \bibitem{Breitenlohner:jf}
1210: P.~Breitenlohner and D.~Z.~Freedman,
1211: ``Stability In Gauged Extended Supergravity,''
1212: Annals Phys.\  {\bf 144} (1982) 249;
1213: %%CITATION = APNYA,144,249;%%
1214: ``Positive Energy In Anti-De Sitter Backgrounds And Gauged Extended 
1215: Supergravity,''
1216: Phys.\ Lett.\ B {\bf 115} (1982) 197.
1217: %%CITATION = PHLTA,B115,197;%%
1218: 
1219: %\cite{Gibbons:aq}
1220: \bibitem{Gibbons:aq}
1221: G.~W.~Gibbons, C.~M.~Hull and N.~P.~Warner,
1222: ``The Stability Of Gauged Supergravity,''
1223: Nucl.\ Phys.\ B {\bf 218} (1983) 173.
1224: %%CITATION = NUPHA,B218,173;%%
1225: 
1226: \bibitem{Lawson89}
1227: B. Lawson and M. Michelsohn, {\it Spin Geometry}, Princeton University Press,
1228: Princeton, NJ (1989).
1229: 
1230: 
1231: \bibitem{Schoen79}
1232: R. Schoen and S. Yau, ``The structure of manifolds with positive scalar
1233: curvature", Manuscripta Math. {\bf 28} (1979) 159;
1234: M. Gromov and H. Lawson, ``Spin and scalar curvature in the presence
1235: of a fundamental group I", Ann of Math {\bf 111} (1980) 209.
1236: 
1237: \bibitem{Hitchen74}
1238: N. Hitchen,  ``Harmonic Spinors", Adv. in Math. {\bf 14} (1974) 1.
1239: 
1240: \bibitem{Stolz90}
1241: S. Stolz, ``Simply Connected Manifolds of Positive Scalar Curvature", 
1242: Bull. Amer. Math. Soc. {\bf 23}, 427 (1990).
1243: 
1244: \bibitem{Lawson80}
1245: M. Gromov and H. Lawson, ``The classification of simply connected
1246: manifolds of positive scalar curvature", Ann of Math {\bf 111} (1980) 423.
1247: 
1248: \bibitem{Dwyer02}
1249: W. Dwyer, T. Schick, and A. Stolz, ``Remarks on a conjecture of 
1250: Gromov and Lawson", math.GT/0208011.
1251: 
1252: \bibitem{Rosenberg94}
1253: J. Rosenberg and S. Stolz, ``A `stable' version of the Gomov-Lawson
1254: conjecture", Contemp. Math {\bf 181} (1995) 405; dg-ga/9407002.
1255: 
1256: %\cite{McInnes:2001xp}
1257: \bibitem{McInnes:2001xp}
1258: B.~T.~McInnes,
1259: ``Topologically induced instability in string theory,''
1260: JHEP {\bf 0103} (2001) 031
1261: [arXiv:hep-th/0101136].
1262: %%CITATION = HEP-TH 0101136;%%
1263: 
1264: \bibitem{Lee87}
1265: R. Schoen, ``Conformal deformation of a  Riemannian metric to constant
1266: scalar curvature", J. Differential Geom. {\bf 20} (1984) 479;
1267: J. Lee and T. Parker, ``The Yamabe Problem", Bull. Am. Math. Soc.
1268: {\bf 17} (1987) 37.
1269: 
1270: 
1271: \bibitem{Witt}
1272: D. Witt, private communication.
1273: 
1274: \bibitem{Pollack}
1275: D. Pollack, ``Nonuniqueness and High Energy Solutions for a Conformally
1276: Invariant Scalar Equation", Comm. Anal. and Geom. {\bf 1} (1993) 347.
1277: 
1278: \bibitem{Pollack2}
1279: D. Pollack, private communication. 
1280: 
1281: \bibitem{SY}
1282: R. Schoen and S. Yau, ``Positivity of the Total Mass of a General Space-Time,''
1283: Phys. Rev. Lett. {\bf 43} (1979) 1457.
1284: 
1285: %\cite{Parker:uy}
1286: \bibitem{Parker:uy}
1287: T.~Parker and C.~H.~Taubes,
1288: ``On Witten's Proof Of The Positive Energy Theorem,''
1289: Commun.\ Math.\ Phys.\  {\bf 84} (1982) 223.
1290: %%CITATION = CMPHA,84,223;%%
1291: 
1292: 
1293: %\cite{Townsend:iu}
1294: \bibitem{Townsend:iu}
1295: P.~K.~Townsend,
1296: ``Positive Energy And The Scalar Potential In Higher Dimensional (Super)Gravity Theories,''
1297: Phys.\ Lett.\ B {\bf 148} (1984) 55.
1298: %%CITATION = PHLTA,B148,55;%%
1299: 
1300: %\cite{DeWolfe:1999cp}
1301: \bibitem{DeWolfe:1999cp}
1302: O.~DeWolfe, D.~Z.~Freedman, S.~S.~Gubser and A.~Karch,
1303: ``Modeling the fifth dimension with scalars and gravity,''
1304: Phys.\ Rev.\ D {\bf 62} (2000) 046008
1305: [arXiv:hep-th/9909134].
1306: %%CITATION = HEP-TH 9909134;%%
1307: 
1308: \bibitem{Hertog03}
1309: T. Hertog, G.T. Horowitz, K. Maeda, to appear.
1310: 
1311: %\cite{Geddes:2001ht}
1312: \bibitem{Geddes:2001ht}
1313: J.~Geddes,
1314: ``The collapse of large extra dimensions,''
1315: Phys.\ Rev.\ D {\bf 65} (2002) 104015
1316: [arXiv:gr-qc/0112026].
1317: %%CITATION = GR-QC 0112026;%%
1318: 
1319: %\cite{Gross:cv}
1320: \bibitem{Gross:cv}
1321: D.~J.~Gross, M.~J.~Perry and L.~G.~Yaffe,
1322: ``Instability Of Flat Space At Finite Temperature,''
1323: Phys.\ Rev.\ D {\bf 25} (1982) 330.
1324: %%CITATION = PHRVA,D25,330;%%
1325: 
1326: %\cite{Banks:1984cw}
1327: \bibitem{Banks:1984cw}
1328: T.~Banks,
1329: ``T C P, Quantum Gravity, The Cosmological Constant And All That..,''
1330: Nucl.\ Phys.\ B {\bf 249} (1985) 332.
1331: %%CITATION = NUPHA,B249,332;%%
1332: 
1333: %\cite{Felder:2002jk}
1334: \bibitem{Felder:2002jk}
1335: G.~N.~Felder, A.~V.~Frolov, L.~Kofman and A.~D.~Linde,
1336: ``Cosmology with negative potentials,''
1337: Phys.\ Rev.\ D {\bf 66} (2002) 023507
1338: [arXiv:hep-th/0202017].
1339: %%CITATION = HEP-TH 0202017;%%
1340: 
1341: %\cite{Kachru:2003aw}
1342: \bibitem{Kachru:2003aw}
1343: S.~Kachru, R.~Kallosh, A.~Linde and S.~P.~Trivedi,
1344: ``De Sitter vacua in string theory,''
1345: arXiv:hep-th/0301240.
1346: %%CITATION = HEP-TH 0301240;%%
1347: 
1348: %\cite{Ford:1999qv}
1349: \bibitem{Ford:1999qv}
1350: L.~H.~Ford and T.~A.~Roman,
1351: ``The quantum interest conjecture,''
1352: Phys.\ Rev.\ D {\bf 60} (1999) 104018
1353: [arXiv:gr-qc/9901074];
1354: %%CITATION = GR-QC 9901074;%%
1355: E.~E.~Flanagan,
1356: ``Quantum inequalities in two dimensional curved spacetimes,''
1357: Phys.\ Rev.\ D {\bf 66} (2002) 104007
1358: [arXiv:gr-qc/0208066].
1359: %%CITATION = GR-QC 0208066;%%
1360: 
1361: %\cite{Kachru:1995wm}
1362: \bibitem{Kachru:1995wm}
1363: S.~Kachru and C.~Vafa,
1364: ``Exact results for N=2 compactifications of heterotic strings,''
1365: Nucl.\ Phys.\ B {\bf 450} (1995) 69
1366: [arXiv:hep-th/9505105].
1367: %%CITATION = HEP-TH 9505105;%%
1368: 
1369: %\cite{Ferrara:1995yx}
1370: \bibitem{Ferrara:1995yx}
1371: S.~Ferrara, J.~A.~Harvey, A.~Strominger and C.~Vafa,
1372: ``Second quantized mirror symmetry,''
1373: Phys.\ Lett.\ B {\bf 361} (1995) 59
1374: [arXiv:hep-th/9505162].
1375: %%CITATION = HEP-TH 9505162;%%
1376: 
1377: 
1378: \bibitem{Penrose02}
1379: R. Penrose, ``On the Instability of Extra Space Dimensions",
1380: in {\em The Future of Theoretical Physics and Cosmology},
1381: Festschrift in honor of S.W. Hawking's 60th birthday, eds. G.W. Gibbons,
1382: S. Rankin, P. Shellard, to appear.
1383: 
1384: \end{thebibliography} 
1385: 
1386: \end{document}
1387: 
1388: 
1389:      
1390:    
1391: 
1392: