1: %\documentclass[aps,amsmath,amssymb,superscriptaddress,letterpaper,preprint]{revtex4}
2: %\documentclass[aps,amsmath,amssymb,superscriptaddress,letterpaper]{revtex4}
3: \documentclass[prd,amsmath,amssymb,groupedaddress,letterpaper,twocolumn]{revtex4}
4: %\documentclass[10pt,letter]{article}
5: \usepackage{amsmath,amssymb,graphicx,bm}
6: %\usepackage{pxfonts}
7: %\usepackage{times}
8: %\usepackage[active]{srcltx}
9:
10:
11:
12: %\setlength{\voffset}{0.25in}
13:
14:
15: \renewcommand{\a}{\alpha}
16: \renewcommand{\b}{\beta}
17: \newcommand{\cA}{{\cal A}}
18: \newcommand{\cB}{{\cal B}}
19: \newcommand{\cC}{{\cal C}}
20: \newcommand{\cH}{{\cal H}}
21:
22: \newcommand{\cK}{{\cal K}}
23: \newcommand{\cL}{{\cal L}}
24: \newcommand{\cM}{{\cal M}}
25: \newcommand{\cN}{{\cal N}}
26: \newcommand{\cV}{{\cal V}}
27: \renewcommand{\d}{\delta}
28: \newcommand{\f}{\phi}
29: \newcommand{\fr}{\tfrac}
30: \newcommand{\nn}{\nonumber}
31: \newcommand{\p}{\partial}
32:
33: \newcommand{\sign}{{\textrm{sign}}}
34: \newcommand{\T}{{\textsf{T}}}
35:
36: \newtheorem{theorem}{Theorem}
37:
38:
39:
40: \begin{document}
41:
42:
43: \title{Instabilities and the null energy condition}
44:
45: \author{Roman~V.~Buniy}
46: \email{roman@uoregon.edu}
47: \author{Stephen~D.H.~Hsu}
48: \email{hsu@duende.uoregon.edu}
49:
50:
51: \affiliation{Institute of Theoretical Science, University of Oregon,
52: Eugene OR 94703-5203}
53:
54:
55:
56:
57:
58:
59:
60:
61:
62: \begin{abstract}
63: We show that violation of the null energy condition implies
64: instability in a broad class of models, including classical gauge
65: theories with scalar and fermionic matter as well as any perfect
66: fluid. When applied to the dark energy, our results imply that $w = p
67: / \rho$ is unlikely to be less than $-1$.
68: \end{abstract}
69: %\pacs{}
70:
71: \maketitle
72:
73:
74:
75: \section{Introduction}\label{I}
76:
77:
78: Energy conditions, or restrictions on the matter energy-momentum
79: tensor $T_{\mu \nu}$, play an important role in general relativity. No
80: classification of the solutions to Einstein's equation is possible
81: without restrictions on $T_{\mu \nu}$, since every spacetime is a
82: solution for some particular choice of energy-momentum tensor. In
83: this letter we demonstrate a direct connection between stability and
84: the null energy condition (NEC) \cite{null}, $T_{\mu \nu} n^\mu n^\nu
85: \ge 0$ for any null vector $n$ (satisfying $g_{\mu\nu}n^\mu n^\nu =
86: 0$). Our main results are: (1) classical solutions of scalar-gauge
87: models which violate the NEC are unstable, (2) a quantum state
88: (including fermions) in which the expectation of the energy-momentum
89: tensor violates the NEC cannot be the ground state, (3) perfect fluids
90: which violate the NEC are unstable. These results suggest that violations of
91: the NEC in physically interesting cases are likely to be only ephemeral.
92:
93: Our results have immediate applications to the dark energy equation of
94: state, often given in terms of $w=p/\rho$. Dark energy has positive
95: energy density $\rho$ and energy-momentum tensor $T_{\mu \nu} = \text
96: {diag\,}(\rho, p, p, p)$ in the comoving cosmological
97: frame. Therefore, $w < -1$ implies violation of the NEC. Instability
98: as a consequence of $w < -1$ was studied previously in scalar
99: models~\cite{instability}.
100:
101: Some results in relativity in which the NEC plays an important role
102: include the classical singularity theorems \cite{Hawking-Ellis},
103: proposed covariant entropy bounds \cite{bousso} and non-existence of
104: Lorentzian wormholes~\cite{wormholes}.
105:
106:
107:
108:
109:
110: \section{Field theories}\label{FT}
111:
112:
113: Consider a theory of scalar, $\phi_a$, and gauge, $A_{a\alpha}$,
114: fields in a fixed $d$-dimensional space-time with the metric
115: $g_{\mu\nu}$. We limit ourselves to theories whose equations of motion
116: are second order differential equations, so the Lagrangian for the
117: system is assumed to depend only on the fields and their first
118: derivatives. We take the Lagrangian density $\cL$ to depend only on
119: the covariant derivative of the field $D_\mu\phi_a$ and the gauge
120: field strength $F_{a\mu\nu}$. The scalars may transform in any
121: representation of the gauge group. We impose Lorentz invariance on
122: $\cL$, but do not require overall gauge invariance. That is, we allow
123: for fixed tensors with gauge indices (but no Lorentz indices) which
124: can be contracted with the fields. For the corresponding action
125: \begin{eqnarray}
126: S=\int
127: d^dx\,|g|^\frac{1}{2}\cL(\phi_a,D_\mu\phi_a,F_{a\mu\nu})\label{FT:S}
128: \end{eqnarray}
129: to be stationary, its first variation has to vanish, $\delta
130: S=0$. This leads to the equations of motion for the fields $\phi_a$ and
131: $A_{a\alpha}$; in the classical analysis we assume that we have found
132: solutions to these equations, about which we expand.
133:
134:
135:
136: \subsection{Null energy condition}\label{S:NEC}
137:
138:
139: The quantities $D_\mu\phi_a$ and $g^{\mu\nu}$ are independent
140: variables. Nevertheless we now prove that there is a relation between the
141: derivatives of $\cL$ with respect to them:
142: %\begin{subequations}
143: \begin{eqnarray}
144: 2\cL_{g^{\mu\nu}} &=& M^{AB}\psi_{A\mu}\psi_{B\nu} + g_{\mu\nu} K,
145: \label{FT:M.g}\\ \cL_{\psi_{A\mu}} &=& M^{AB}{\psi_B}^\mu +
146: \epsilon^{\mu \nu_2 \ldots \nu_{d}} {L^A}_{\nu_2 \ldots \nu_{d}}.
147: \label{FT:M.psi}
148: \end{eqnarray}
149: %\label{FT:M}
150: %\end{subequations}
151: In our notation $\psi_{A\mu}=(D_\mu\phi_a,F_{a\alpha\mu})$, where the
152: abstract index $A$ may run over both Lorentz and color indices, as
153: well as the type of field. So, $\psi_{A\mu}$ is a list of objects,
154: each of which has a Lorentz index $\mu$. The value of $A$ specifies an
155: element of this list.
156:
157: The relations are obtained
158: by noting that for each and every $g^{\mu\nu}$ in $\cL$ there are two
159: $\psi$s attached to it, except for the curved space totally
160: antisymmetric tensor $\vert g \vert^{-\frac{1}{2}} \epsilon^{\nu_1
161: \ldots \nu_d}$, which gives rise to the $K$ term in
162: Eq.~(\ref{FT:M.g}). Similarly, differentiation with respect to
163: $\psi_{A\mu}$ yields the $M$ and $L$ terms in
164: Eq.~(\ref{FT:M.psi}).
165:
166: Figure~\ref{figure} represents the most general Lagrangian of type
167: considered in this paper. Each dot represents a Lorentz index and a
168: line connecting them denotes contraction using the metric. Rectangles
169: (with two indices) are field strengths, small circles covariant
170: derivatives of scalar fields, and a large circle an epsilon
171: tensor. Finally, the block $X$ represents the remainder of the
172: diagram. Because the product of two epsilon tensors can be rewritten
173: as a sum of products of metric tensors $g$, we need to consider only
174: figures with at most one epsilon tensor, and therefore can assume that
175: $X$ contains none. (For generality, we include an epsilon tensor in
176: the figure, although of course $\cL$ need not contain a
177: parity-violating component.) All Lorentz indices are ultimately
178: contracted, and we suppress color indices for simplicity.
179:
180: \begin{figure}[h!]
181: \includegraphics[width=6cm]{figure-1}
182: \caption{Representation of the most general Lagrangian of the type
183: considered in this paper. Each dot represents a Lorentz index, and a
184: line connecting them denotes contraction using the metric. Rectangles
185: (with two indices) are field strengths, small circles covariant
186: derivatives of scalar fields, and large circles epsilon
187: tensors. Finally, the block $X$ represents the remainder of the
188: diagram. All Lorentz indices are ultimately contracted, and we
189: suppress color indices for simplicity. In graphical terms, $M$ is
190: obtained by simply removing the shaded elements.}
191: \label{figure}
192: \end{figure}
193:
194:
195: Consider the labelled portion of the figure, which equals
196: \begin{equation}
197: \cL = \vert g \vert^{-\frac{1}{2}} \epsilon^{\alpha \ldots}
198: F_{\alpha \mu} \, g^{\mu \nu} F_{\nu \beta} \, g^{\beta \gamma}
199: X_{\gamma \ldots}.
200: \label{FT:M.L}
201: \end{equation}
202: Again, we suppress color indices for simplicity, as they do not affect
203: the proof. We include an epsilon tensor in the analysis, although
204: $\cL$ may or may not contain one (in the parity-preserving case the
205: epsilon tensor in Eq.~(\ref{FT:M.L}) is replaced by the metric). By
206: differentiation of the indicated portion, we obtain
207: Eqs.~(\ref{FT:M.g}) and (\ref{FT:M.psi}) with
208: \begin{eqnarray}
209: M^{\alpha \beta} &=& - \vert g \vert^{-\frac{1}{2}} \left(
210: \epsilon^{\alpha \ldots} g^{\beta \gamma} + \epsilon^{\beta \ldots}
211: g^{\alpha \gamma} \right) X_{\gamma \ldots}, \\ K &=& \vert g
212: \vert^{-\frac{1}{2}} \epsilon^{\alpha \ldots} F_{\alpha \rho} \,
213: g^{\rho \sigma} F_{\sigma \beta} \, g^{\beta \gamma} X_{\gamma \ldots}, \\
214: {L^\alpha}_{\ldots} &=& - \vert g \vert^{-\frac{1}{2}} g^{\alpha \rho}
215: F_{\rho \beta} \,g^{\beta \gamma} X_{\gamma \ldots}. \label{FT:L}
216: \end{eqnarray}
217: Note the derivative $\cL_{F_{\nu \beta}}$ generates a contribution to
218: $M$ which is matched by a corresponding contribution from
219: $2 \cL_{g^{\beta \gamma}}$. Other contractions of fields with $g_{\mu
220: \nu}$ (i.e., as indicated in the figure) can be analyzed similarly.
221: In graphical terms, $M$ can be obtained from the figure for $\cL$ by simply
222: removing two $\psi$s, in this case the shaded elements of the figure.
223:
224:
225:
226:
227:
228:
229:
230: For the energy-momentum tensor which couples to gravity,
231: \begin{eqnarray}
232: T_{\mu\nu} =-\cL g_{\mu\nu} +2\cL_{g^{\mu\nu}}, \label{FT:T}
233: \end{eqnarray}
234: the NEC then requires
235: \begin{eqnarray}
236: \Psi_A M^{AB}\Psi_B\ge 0, \label{FT:NEC.M}
237: \end{eqnarray}
238: where $\Psi_A=\psi_{A\mu}n^\mu$. Thus, to satisfy the NEC, $M^{AB}$ has to
239: be positive semidefinite. This property is crucial for stability of
240: solutions, to which we now turn.
241:
242:
243:
244:
245: \subsection{Stability}\label{S:Stability}
246:
247:
248: To study the stability of the solution $\psi_A(x)$, we consider the
249: second variation of the Lagrangian,
250: \begin{eqnarray}
251: \d^2 \cL &=& \cL_{\psi_A\psi_B}
252: \delta\psi_A\delta\psi_B +
253: 2\cL_{\psi_{A}\psi_{B;\lambda}}\delta\psi_{A}\delta\psi_{B;\lambda}
254: \nn \\ &+& \cL_{\psi_{A;\mu}\psi_{B;\nu}}
255: \delta\psi_{A;\mu}\delta\psi_{B;\nu}.\label{FT:d2L}
256: \end{eqnarray}
257: Here quantities $\cL_{\psi_A}=\p\cL/\p\psi_A$, etc. are evaluated at
258: $\psi_A(x)$. Also notice that
259: $\psi_{A;\mu}=(D_\mu\phi_a,A_{a\alpha;\mu})$, the covariant
260: derivatives of $\psi_A$, are different from
261: $\psi_{A\mu}=(D_\mu\phi_a,F_{a\alpha\mu})$.
262:
263: Let us use a locally inertial frame in which the metric is reduced to
264: $\bar{g}_{\mu\nu} =\text{diag\,}(1,-1,\ldots,-1)$; all quantities in
265: this frame are designated by a bar. For the
266: Lagrangian~(\ref{FT:d2L}), the canonical momentum is
267: \begin{eqnarray}
268: \delta\bar{\pi}^A =
269: 2\cL_{\bar\psi_{B}\bar\psi_{A;0}}\delta\bar\psi_{B} +
270: 2 \cL_{\bar\psi_{A;0}\bar\psi_{B;\nu}} \delta\bar\psi_{B;\nu}
271: \label{FT:dpi}
272: \end{eqnarray}
273: which leads to the following effective Hamiltonian for fluctuations
274: about the classical solution:
275: \begin{eqnarray}
276: \delta^2\cH &=& -\cL_{\bar\psi_{A}\bar\psi_{B}} \delta\bar\psi_{A}
277: \delta\bar\psi_{B} -2\cL_{\bar\psi_{A}\bar\psi_{B;j}}
278: \delta\bar\psi_{A} \delta\bar\psi_{B;j} \label{FT:d2H} \\ &+&
279: \cL_{\bar\psi_{A;0}\bar\psi_{B;0}} \delta\bar\psi_{A;0}
280: \delta\bar\psi_{B;0} - \cL_{\bar\psi_{A;i}\bar\psi_{B;j}}
281: \delta\bar\psi_{A;i} \delta\bar\psi_{B;j}.\nn
282: \end{eqnarray}
283: Here $\delta\bar{\psi}_{A;0}$ are functions of $\delta\bar{\pi}^B$,
284: $\delta\bar{\psi}_{B}$ and $\delta\bar{\psi}_{B;i}$ as found from
285: Eq.~(\ref{FT:dpi}). The first term on the right hand side of
286: Eq.~(\ref{FT:d2H}) is a potential term, which we denote by
287: $\delta^2\cV$, and the last two terms are kinetic terms, denoted
288: $\delta^2\cK$.
289:
290: If the kinetic energy $\delta^2\cK$ is
291: negative, then the system described by the Hamiltonian $\delta^2\cH
292: =\delta^2\cK +\delta^2\cV$ is (locally) unstable.
293: If $\delta^2\cV$ is positive then small
294: perturbations will cause the classical solutions to grow
295: exponentially away from the original stationary point. However, it is
296: possible to have classical stability if one chooses $\delta^2\cV$ to
297: be negative; in this case we have an upside-down potential with
298: negative kinetic term, or a ``phantom''. Such models necessarily
299: exhibit quantum instabilities \cite{phantom}.
300: Notice, the second term in Eq.~(\ref{FT:d2H}) is linear in the
301: fluctuations and their derivatives, and therefore can never
302: stabilize the system.
303:
304: To investigate the kinetic terms, we calculate second derivatives of
305: $\cL$ needed in Eq.~(\ref{FT:d2H}). Using Eq.~(\ref{FT:M.psi}) we
306: obtain
307: \begin{equation}
308: \cL_{\psi_{A\mu}\psi_{B\nu}} = M^{AB}g^{\mu\nu} +N^{A\mu
309: B\nu}.\label{FT:MN}
310: \end{equation}
311: The separation of the second derivative into the first and second
312: terms in Eq.~(\ref{FT:MN}) is natural: $g^{\mu\nu}$ appears only in
313: the first term, and $N$ represents all remaining terms. $N$ contains
314: terms obtained by differentiating $M$ and $L$ with respect to
315: $\psi_{B\nu}$, plus additional terms if $\psi$ is a field
316: strength. The $\nu$ index obtained from these $\psi_{B\nu}$
317: derivatives is attached to a field and not the metric $g^{\mu
318: \nu}$. (Also, $L$ does not contain an epsilon tensor since $X$ does not.)
319: Finally, notice that even though $\psi_{A\mu}$ and $\psi_{A;\mu}$ differ,
320: the derivatives of $\cL$ with respect to them coincide due to the form
321: of the action~(\ref{FT:S}). Thus the kinetic term becomes
322: \begin{eqnarray}
323: \delta^2\cK &=& \left(\bar{M}^{AB}+\bar{N}^{A0B0}\right)
324: \delta\bar\psi_{A;0}\delta\bar\psi_{B;0}\nn\\ &+&
325: \left(\bar{M}^{AB}\delta^{ij} -\bar{N}^{AiBj}\right)
326: \delta\bar\psi_{A;i}\delta\bar\psi_{B;j}.\label{FT:d2K}
327: \end{eqnarray}
328:
329:
330: We now prove that nonnegativeness of the kinetic term $\delta^2\cK$
331: implies positive semidefiniteness of the matrix $M$. Indeed,
332: suppose that $M$ is not positive semidefinite. In such case, the
333: matrix $M$ has at least one negative eigenvalue, which means that
334: there is a basis in which the matrix is diagonal with at least one
335: negative entry,
336: $\tilde{M}=\text{diag\,}(\tilde{m}_1,\ldots,\tilde{m}_n)$
337: ($\tilde{m}_1<0$). (Quantities in this basis are designated with a
338: tilde.) Let us choose such field variations that are nonzero
339: only in the direction of the negative eigenvalue:
340: $\delta\tilde{\psi}_{1;\mu}\not =0$ and $\delta\tilde{\psi}_{A;\mu}=0$
341: $(A>1)$. We further restrict $d-1$ quantities
342: $\delta\tilde{\psi}_{1;i}$ to satisfy the following equation:
343: \begin{eqnarray}
344: \tilde{N}^{1010}\delta\tilde{\psi}_{1;0} \delta\tilde{\psi}_{1;0}
345: =\tilde{N}^{1i1j}\delta\tilde{\psi}_{1;i}\delta\tilde{\psi}_{1;j}.
346: \label{FT:Ncondition}
347: \end{eqnarray}
348: These conditions make the kinetic term of Eq.~(\ref{FT:d2K}) negative,
349: \begin{eqnarray}
350: \delta^2\cK =\tilde{m}_1\left[(\delta\tilde{\psi}_{1;0})^2
351: +\sum_i(\delta\tilde{\psi}_{1;i})^2\right] <0,
352: \label{FT:d2K1}
353: \end{eqnarray}
354: thus proving that in order for $\delta^2\cK$ to be nonnegative, the
355: matrix $M$ has to be positive semidefinite.
356:
357: Using the result established in the previous paragraph, we conclude
358: that solutions to the theory given by the action~(\ref{FT:S}) are
359: stable only if the matrix $M$ is positive semidefinite.
360:
361: Combining the relations between nonnegativeness of $\delta^2\cK$ and
362: positive semidefiniteness of $M$ on one hand, and the NEC and
363: positive semidefiniteness of $M$ on the the other hand, we
364: conclude that for the theory given by the action~(\ref{FT:S}), only
365: solutions satisfying the NEC can be stable.
366:
367: We can deduce similar results for quantum systems. Suppose there
368: exists a quantum state $\vert \alpha \rangle$ and a null vector $n^\mu$
369: such that
370: \begin{equation}
371: \langle \alpha \vert T_{\mu\nu} \vert \alpha \rangle n^{\mu} n^{\nu} =
372: \langle \alpha \vert M^{AB} \Psi_{A} \Psi_{B} \vert \alpha \rangle <
373: 0,
374: \end{equation}
375: so that the NEC is violated in a quantum averaged sense. Define a
376: basis $\vert \phi \rangle$ in which the operator ${\cal M} = M^{AB}
377: \Psi_A \Psi_B$ is diagonal:
378: \begin{equation}
379: {\cal M} \vert \phi \rangle = m(\phi) \vert \phi \rangle.
380: \end{equation}
381: Then violation of the quantum averaged NEC implies
382: \begin{equation}
383: \sum_{\phi \phi'} \, \langle \alpha \vert \phi \rangle \langle \phi
384: \vert {\cal M} \vert \phi' \rangle \langle \phi' \vert \alpha \rangle
385: \\ = \sum_{\phi} \, \vert \langle \alpha \vert \phi \rangle \vert^2
386: \,m(\phi) < 0.
387: \end{equation}
388: This means that there exist eigenstates $\vert \phi \rangle$, whose
389: overlap with $| \alpha \rangle$ is non-zero and on which the operator
390: ${\cal M}$ has negative eigenvalues. This requires that $M$ and
391: hence $\delta^2\cK$ is not positive semidefinite; by continuity, this
392: must also be the case in a ball $B$ in the Hilbert space of $\vert \phi
393: \rangle$.
394:
395: As a further consequence, we can conclude that a state $\vert \alpha
396: \rangle$ in which the NEC is violated cannot be the ground
397: state~\footnote{In some curved spacetimes there may not be a well-defined
398: ground state. In de Sitter space, quantum corrections can cause
399: violation of the NEC: V.~K.~Onemli and R.~P.~Woodard, Class.\ Quant.\
400: Grav. {\bf 19}, 4607 (2002); Phys.\ Rev.\ D {\bf 70}, 107301
401: (2004).}. Suppose that $\vert \alpha \rangle$ is an energy eigenstate:
402: $H \vert \alpha \rangle = E_\alpha \vert \alpha \rangle$. An
403: elementary result from quantum mechanics is that $\vert \alpha
404: \rangle$ can be the ground state only if
405: \begin{equation}
406: E_\alpha = \langle \alpha \vert H \vert \alpha \rangle \leq \langle
407: \alpha' \vert H \vert \alpha' \rangle
408: \end{equation}
409: for all normalized states $\vert \alpha' \rangle$ which need not be
410: energy eigenstates. However, it is possible to reduce the expectation
411: value of $H$ by perturbing $\vert \alpha \rangle$. Specifically, we
412: adjust $\vert \alpha \rangle$ only in the ball $B$, where we know from
413: Eqs.~(\ref{FT:d2K})--(\ref{FT:d2K1}) that there are perturbations
414: which reduce the expectation of the kinetic energy without changing
415: the expectation of the potential.
416:
417: Note that the discussion above is in terms of unrenormalized (bare)
418: quantities. The renormalized expectation $\langle \alpha \vert {\cal
419: M}_{\rm ren} \vert \alpha \rangle = \langle \alpha \vert {\cal M}
420: \vert \alpha \rangle - \langle 0 \vert {\cal M} \vert 0 \rangle$
421: (where $\vert 0 \rangle$ is the flat-space QFT ground state) could be
422: negative (e.g., as in the Casimir effect \cite{null}), but this is
423: possible only if $\vert \alpha \rangle$ is not $\vert 0 \rangle$.
424:
425: In known cases of NEC violation, such as the Casimir vacuum or black
426: hole spacetime, it is only the {\it renormalized} energy-momentum
427: tensor which violates the NEC. As a simple example, consider a real
428: scalar field $\phi$. The energy-momentum tensor is simply $T_{\mu \nu}
429: = \partial_\mu \phi \, \partial_\nu \phi$ plus terms proportional to
430: $g_{\mu\nu}$ which do not play a role in the NEC. Then, ${\cal M} = (
431: n^{\mu} \partial_{\mu} \phi )^2$ is a Hermitian operator with positive
432: eigenvalues. Therefore, its expectation value in {\it any} state is
433: positive: $\langle \alpha \vert {\cal M} \vert \alpha \rangle > 0,$
434: for any $\vert \alpha \rangle$, including the Hartle-Hawking, Casimir
435: or flat-space vacuum. We can verify this by direct calculation,
436: computing the energy-momentum tensor using point-splitting
437: regularization:
438: \begin{eqnarray}
439: \langle 0 \vert T_{\mu\nu} (x,x') \vert 0 \rangle n^\mu n^\nu = \frac{2
440: [n^\mu (x-x')_\mu]^2 }{\pi^2 \vert x-x'\vert^6},
441: \end{eqnarray}
442: which is manifestly positive. Note that $\langle \alpha \vert {\cal M}
443: \vert \alpha \rangle > 0$ for all $\vert \alpha \rangle$, since the
444: bare expectation is always dominated by the UV contribution. Now, had
445: we taken a {\it negative} kinetic energy term for the scalar, the
446: overall sign of ${\cal M}$ would change, allowing violation of the
447: NEC. But, this model is clearly unstable, in accordance with our
448: results.
449:
450:
451:
452:
453: \subsection{Fermions}\label{FF}
454:
455:
456: To this point we have only considered bosonic fields. We now extend
457: our analysis to systems with fermions, adding to our Lagrangian the
458: term
459: \begin{equation}
460: \cL^{(\text{f})} = \bar{\psi} ( i D\!\!\!\!/ - m ) \psi.
461: \end{equation}
462: (A scalar-fermion coupling can be treated similarly, as can a Weyl
463: fermion, whose determinant is the square root of the Dirac
464: determinant.) Then, for any fixed gauge field background the fermions
465: can be integrated out directly in favor of a non-local correction to
466: the action for bosonic fields
467: \begin{equation}
468: - \sum_{\lambda_l > 0} \ln ( \lambda_l^2 + m^2 ),
469: \end{equation}
470: where $D\!\!\!\!/ \, \psi_l = \lambda_l \psi_l$ is the eigenvalue
471: equation for the Dirac operator, with $\lambda_l$ real. This shifts
472: the energy-momentum tensor by
473:
474: \begin{equation}
475: T_{\mu \nu}^{(\text{f})} = -2\vert g \vert ^{-\frac{1}{2}}
476: \sum_{\lambda_l > 0} \frac{1}{\lambda_l^2 + m^2} \frac{\d
477: \lambda_l^2}{\d g^{\mu \nu}}.
478: \end{equation}
479: Now write $\lambda_l^2 \psi_l = {D\!\!\!\!/ }^{\,2} \psi_l = (g^{\mu
480: \nu} D_\mu D_\nu -\tfrac{1}{2}i \sigma^{\mu\nu} F_{\mu\nu} )\psi_l $
481: and use the orthonormality of the eigenfunctions to obtain
482: \begin{equation}
483: \lambda_l^2 = \int d^dx \, |g|^\frac{1}{2} \, g^{\mu \nu}
484: \,\psi^\dagger_l D_\mu D_\nu \psi_l ~+~ \ldots,
485: \end{equation}
486: where the ellipsis denote terms which do not contain $g^{\mu\nu}$.
487: After integration by parts, the contribution to the NEC from fermions
488: is then
489: \begin{equation}
490: T^{(\text{f})}_{\mu\nu} n^\mu n^\nu = \sum_{\lambda_l >
491: 0} \frac{2} {\lambda_l^2 + m^2} \,
492: ( n \cdot D \psi_l)^\dagger ( n \cdot D \psi_l ).
493: \end{equation}
494: This additional contribution is always positive. So, the conclusions
495: of the previous section are unmodified by the presence of fermions:
496: violation of the NEC implies the bosonic kinetic energy is not
497: positive semidefinite.
498:
499:
500:
501: \section{Perfect fluid}\label{F}
502:
503:
504: A macroscopic system may be approximately described as a perfect fluid
505: if the mean free path of its components is small compared to the
506: length scale of interest. For the dark energy, this length scale is of
507: cosmological size. A perfect fluid is described by the
508: energy-momentum tensor
509: \begin{eqnarray}
510: T_{\mu\nu}=(\rho+p)u_\mu u_\nu-pg_{\mu\nu},\label{F:T}
511: \end{eqnarray}
512: where $\rho$ and $p$ are the energy density and pressure of the fluid
513: in its rest frame, and $u_\mu$ is its velocity. Let $j^\mu=J u^\mu$ be
514: the conserved current vector (${j^\mu}_{;\mu}=0$), and $J=(j_\mu
515: j^\mu)^\frac{1}{2}$ the particle density.
516:
517: The energy-momentum can be written \cite{Hawking-Ellis,Jackiw:2004nm}
518: as
519: \begin{eqnarray}
520: T_{\mu\nu}=(f-J f')g_{\mu\nu}+(f'/J) j_\mu j_\nu,\label{F:T1}
521: \end{eqnarray}
522: where, comparing with Eq.~(\ref{F:T}), we have $\rho = f(J)$ and $p =
523: Jf' - f$. The function $f(J)$ implicitly determines the equation of state.
524:
525: The NEC for the tensor~(\ref{F:T1}) becomes
526: \begin{eqnarray}
527: T_{\mu\nu}n^\mu n^\nu=(f'/J)(j_\mu n^\mu)^2\ge 0.\label{F:NEC}
528: \end{eqnarray}
529: Thus, perfect fluids with negative $f'(J)$ violate the NEC. Below, we
530: demonstrate that $f'(J) < 0$ implies an instability.
531:
532: Recall that the speed of sound in a fluid is given by $s = (dp /
533: d\rho)^{\frac{1}{2}} = (J f'' / f')^{\frac{1}{2}}$, and that complex
534: $s$ implies an instability. Note $f'(J)$ cannot change its sign
535: without producing an instability. Indeed, if it were to change its
536: sign at some $J_*$, then $s$ would be complex for either $J$ larger or
537: smaller than $J_*$, depending on the sign of $f''(J_*)$. ($f''$
538: cannot also change sign at $J_*$.) Therefore, if $f'$ is negative
539: anywhere, then it is negative everywhere, and to avoid complex $s$,
540: $f''$ must be negative everywhere.
541:
542: However, if $f'$ and $f''$ are everywhere negative, then the fluid is
543: unstable with respect to clumping. To see this, we first deduce the
544: dependence of the fluid free energy $F$ on particle number $N =
545: JV$. Note that $( \partial F / \partial V )\vert_{T,N} = - p = N
546: \partial (f/J) / \partial V$. By integration we find $F = N [ f/J -
547: h(T)]$, where the first term is just the energy $E$ and $N h(T) = T
548: S$, where $S$ is the entropy. It is easy to see that $(\partial^2 F
549: /\partial J^2 )\vert_{T,V} = V f''$.
550:
551: Now consider two adjacent regions of the fluid
552: with identical volumes. Suppose we transfer a small amount of matter
553: $\delta J$ from one volume to another; the resulting change in total
554: free energy is given by $\tfrac{1}{2} V f''(J) (\delta J)^2 <
555: 0$. We see that the system can decrease its free energy by clumping into
556: over- and under-dense regions. This itself is an instability, which
557: results in a runaway to infinitely negative free energy unless the
558: assumption of negative $f'$ (violation of NEC) or negative $f''$ (real
559: $s$) ceases to hold.
560:
561: \bigskip
562:
563: \begin{center}
564: \textbf{Acknowledgments}
565: \end{center}
566:
567: The authors thank R.~Bousso, A.~Jenkins, B.~Murray, M.~Schwartz,
568: D.~Soper and M.~Wise for useful comments. This work was supported by
569: the Department of Energy under DE-FG06-85ER40224.
570:
571:
572:
573:
574:
575:
576:
577:
578:
579:
580:
581:
582:
583:
584:
585:
586: \begin{thebibliography}{99}
587:
588:
589: \bibitem{null}
590: For discussion of the NEC and possible violations, see, e.g.,
591: M.~Visser,
592: \emph{Lorentzian wormholes: From Einstein to Hawking}, AIP, Woodbury USA,
593: 1995; E.~E.~Flanagan and R.~M.~Wald,
594: %``Does backreaction enforce the averaged null energy condition in
595: %semiclassical gravity?,''
596: Phys.\ Rev.\ D {\bf 54}, 6233 (1996);
597: %[arXiv:gr-qc/9602052];
598: %%CITATION = GR-QC 9602052;%%
599: J.~Fewster and T.~A.~Roman,
600: %``Null energy conditions in quantum field theory,''
601: Phys.\ Rev.\ D {\bf 67}, 044003 (2003);
602: %[arXiv:gr-qc/0209036];
603: L.~H.~Ford,
604: %``The classical singularity theorems and their quantum loopholes,''
605: Int.\ J.\ Theor.\ Phys.\ {\bf 42}, 1219 (2003).
606: %[arXiv:gr-qc/0301045].
607: %%CITATION = GR-QC 0301045;%%
608:
609: \bibitem{instability}
610: S.~D.~H.~Hsu, A.~Jenkins and M.~B.~Wise,
611: %``Gradient instability for w<-1,''
612: Phys.\ Lett.\ B {\bf 597}, 270 (2004).
613: %[arXiv:astro-ph/0406043].
614: %%CITATION = ASTRO-PH 0406043;%%
615: A.~Vikman, Phys.\ Rev.\ D {\bf 71}, 023515 (2005).
616:
617:
618: \bibitem{Hawking-Ellis} S.~W.~Hawking and G.~F.~R.~Ellis, \emph{The
619: Large Scale Structure of Space-time}, Cambridge University Press,
620: Cambridge, 1973.
621:
622: \bibitem{bousso}
623: R.~Bousso,
624: %``The holographic principle,''
625: Rev.\ Mod.\ Phys.\ {\bf 74}, 825 (2002).
626: %[arXiv:hep-th/0203101].
627: %%CITATION = HEP-TH 0203101;%%
628:
629: \bibitem{wormholes} M.~S.~Morris and K.~S.~Thorne,
630: %``Wormholes In Space-Time And Their Use For Interstellar Travel: A Tool For
631: %Teaching General Relativity,''
632: Am.\ J.\ Phys.\ {\bf 56}, 395 (1988); J.~L.~Friedman, K.~Schleich and
633: D.~M.~Witt,
634: %``Topological censorship,''
635: Phys.\ Rev.\ Lett.\ {\bf 71}, 1486 (1993) [Erratum-ibid.\ {\bf 75},
636: 1872 (1995)];
637: %[arXiv:gr-qc/9305017];
638: D.~Hochberg and M.~Visser,
639: %``Geometric structure of the generic static traversable wormhole throat,''
640: Phys.\ Rev.\ D {\bf 56}, 4745 (1997);
641: %[arXiv:gr-qc/9704082];
642: M.~Visser and D.~Hochberg,
643: %``Generic wormhole throats,''
644: arXiv:gr-qc/9710001.
645:
646: \bibitem{phantom}
647: R.~R.~Caldwell,
648: %``A Phantom Menace?,''
649: Phys.\ Lett.\ B {\bf 545}, 23 (2002);
650: %[arXiv:astro-ph/9908168];
651: S.~M.~Carroll, M.~Hoffman and M.~Trodden,
652: %``Can the dark energy equation-of-state parameter w be less than -1?,''
653: Phys.\ Rev.\ D {\bf 68}, 023509 (2003).
654: %[arXiv:astro-ph/0301273].
655:
656: %\cite{Jackiw:2004nm}
657: \bibitem{Jackiw:2004nm} R.~Jackiw, V.~P.~Nair, S.~Y.~Pi and
658: A.~P.~Polychronakos,
659: %``Perfect fluid theory and its extensions,''
660: J.\ Phys.\ A {\bf 37}, R327 (2004).
661: %[arXiv:hep-ph/0407101].
662: %%CITATION = HEP-PH 0407101;%%
663:
664:
665:
666: \end{thebibliography}
667:
668:
669:
670:
671:
672:
673:
674:
675:
676:
677: \end{document}