hep-ph0201082/LV.tex
1: % ****** Start of file apssamp.tex ******
2: %
3: %   This file is part of the APS files in the REVTeX 3.0 distribution.
4: %   Version 3.0 of REVTeX, November 10, 1992.
5: %
6: %   Copyright (c) 1992 The American Physical Society.
7: %
8: %   See the REVTeX 3.0 README file for restrictions and more information.
9: %
10: %
11: %
12: % \documentstyle[preprint,eqsecnum,aps]{revtex}
13: %\documentstyle[twocolumn,aps,epsf]{revtex}
14: 
15: \def\btt#1{{\tt$\backslash$#1}}
16: %\begin{document}
17: %\draft
18: 
19: \hyphenation{di-men-sion-al}
20: \def \half {{1\over 2}}
21: \newcommand{\sss}{{\scriptscriptstyle }}
22: \newcommand{\sperp}{{\scriptscriptstyle\perp}}
23: \def\Square{{\vbox {\hrule height 0.6pt\hbox{\vrule width 0.6pt\hskip 3pt
24:         \vbox{\vskip 6pt}\hskip 3pt \vrule width 0.6pt}\hrule height 0.6pt}}}
25: \def\eV{\hbox{eV}}
26: \def\TeV{\hbox{TeV}}
27: \def\roughly#1{\mathrel{\raise.3ex\hbox{$#1$\kern-.75em
28: \lower1ex\hbox{$\sim$}}}}
29: \def\lsim{\roughly<}
30: \def\gsim{\roughly>}
31: \def\pref#1{(\ref{#1})}
32: \def\nn{\nonumber}
33: \def\eq{\begin{equation}}
34: \def\eeq{\end{equation}}
35: \def\eqa{\begin{eqnarray}}
36: \def\eeqa{\end{eqnarray}}
37: \def\cG{{\cal G}}
38: \def\bk{{\bf k}}
39: \def\bp{{\bf p}}
40: \def\scA{{\cal A}}
41: \def\scB{{\cal B}}
42: \def\scC{{\cal C}}
43: \def\scD{{\cal D}}
44: \def\scE{{\cal E}}
45: \def\eps{\epsilon}
46: \def\veps{\varepsilon}
47: \def\dsl{\hbox{/\kern-.6600em$D$}}
48: \def\psl{\hbox{/\kern-.6600em$P$}}
49: \def\qsl{\hbox{/\kern-.5600em$Q$}}
50: \def\ksl{\hbox{/\kern-.5600em$k$}}
51: \def\g{\gamma}
52: \def\sss{\scriptscriptstyle}
53: \def\a{\alpha}
54: \def\phm{{\phantom{-}\,}}
55: \def\phl{\; {\phantom{\cal L}}}
56: 
57: %% Please only comment out this line
58: \documentstyle[prl,aps,epsf]{revtex}
59: \begin{document}
60: \twocolumn[\hsize\textwidth\columnwidth\hsize\csname@twocolumnfalse%
61: \endcsname
62: \rightline{McGill-01/26, UW/PT 01-27, {\tt hep-th/0201082}} \vspace{3mm}
63: 
64: \draft
65: \title{Loop-Generated Bounds on Changes to the Graviton Dispersion Relation}
66: 
67: \author{C.P.~Burgess,${}^{a}$ J. Cline,${}^a$, E. Filotas${}^a$,
68: J. Matias,${}^b$ and G.D. Moore${}^c$}
69: 
70: \address{${}^a$ Physics Department, McGill University,
71: 3600 University Street, Montr\'eal, Qu\'ebec, Canada H3A 2T8.\\
72: ${}^b$ IFAE, Universitat Aut\`onoma de Barcelona, Spain.\\
73: ${}^c$ Department of Physics, University of Washington, Seattle, WA 98195, USA.}
74: %
75: \maketitle
76: 
77: \begin{abstract}
78: {We identify the effective theory appropriate to the propagation
79: of massless bulk fields in brane-world scenarios, to show that the
80: dominant low-energy effect of asymmetric warping in the bulk is to
81: modify the dispersion relation of the effective 4-dimensional
82: modes. We show how such changes to the graviton dispersion
83: relation may be bounded through the effects they imply, through
84: loops, for the propagation of standard model particles. We compute
85: these bounds and show that they provide, in some cases, the
86: strongest constraints on nonstandard gravitational dispersions.
87: The bounds obtained in this way are the strongest for the fewest
88: extra dimensions and when the extra-dimensional Planck mass is the
89: smallest. Although the best bounds come for warped 5-D scenarios,
90: for which $M_5$ is $O(\hbox{TeV})$, even in 4 dimensions the
91: graviton loop can lead to a bound on the graviton speed which is
92: comparable with other constraints.}
93: \end{abstract}
94: %
95: %
96: \pacs{PACS numbers:
97: 11.10.Kk, 04.50.+h}
98: ]
99: 
100: %
101: %
102: %%%%%%%%%%%\narrowtext  %<- CLH to get two columns a la PRD
103: % aim for around 700 lines
104: 
105: \section{Introduction}
106: 
107: Although physics in more than the traditional four dimensions has been
108: long speculated to be important for describing phenomena at energies
109: above the electroweak scale, the idea is presently enjoying additional
110: scrutiny because of two more recent developments.
111: 
112: First came the
113: understanding of the strong-coupling limit of string theory in terms
114: of 11-dimensional supergravity interacting on spaces with boundaries
115: \cite{HW}. Together with the realization that observable particles
116: can be trapped on these boundaries, or on D-branes, this understanding
117: freed the string scale from the Planck mass, allowing it to be
118: as low as the weak scale \cite{ADD} or at intermediate or
119: grand-unified scales \cite{OtherScales}. Such low string scales
120: allow extra dimensions to be comparatively large, and so (potentially)
121: to have much richer implications at experimentally accessible energies.
122: 
123: Second came the observation that even very small extra dimensions
124: might also have interesting low-energy implications if their geometries
125: are `warped' \cite{RSI,RSII}. In this framework, the four-dimensional metric
126: has a nontrivial dependence on position in the extra dimensions,
127: allowing four-dimensional properties like masses and couplings to
128: depend in an interesting way on an observer's position within the
129: extra dimensions.
130: 
131: Although not required by either approach, the broader class of
132: geometries which are allowed once observable particles are
133: confined to a brane includes many configurations which violate
134: Lorentz invariance within the observable four dimensions by
135: picking out a preferred frame \cite{Lorentz,DRV}, such as by
136: having gravitating objects displaced from our brane in the extra
137: dimensions. For example, within the warped framework one can have
138: a 5D line element of the form
139: %
140: \eq \label{ads}
141:     ds^2 = -(\a+\Delta) dt^2 + \a d\vec x^{\,2} + {dr^2\over
142:     \a+\Delta}
143: \eeq
144: %
145: where $t,$ $\vec{x}$ are 4 dimensional time and space, $r$ is the
146: extra dimension, $\a(r) = r^2/l^2$ and $\Delta(r) = - \mu/r^2 +
147: Q^2/r^4$. If $\mu=Q=0$ this is anti-de Sitter space with curvature
148: radius $l$, and there is no Lorentz violation.  But if $\mu\neq
149: 0$, it is the AdS-Reissner-N\"ordstrom metric with a singularity
150: at $r=0$ and a speed of light which varies with $r$ as $c^2(r)=1+
151: {\Delta\over\a}$. There is a preferred frame with 4-velocity
152: $u^\mu = (1,0,0,0)$ in these coordinates.  In this example, the
153: preferred frame is not due to visible matter, but rather the
154: presence of a ``black brane'' at $r=0$ which is displaced from our
155: brane in the extra dimension. {\it A priori} its effects need not
156: be small, and could cause observable phenomena.
157: 
158: Very strong limits exist on the size of various Lorentz-violating
159: effects involving ordinary particles \cite{LVBounds,Glashow}, so
160: one expects these to provide the most stringent tests of any
161: Lorentz-violation effects predicted in the brane-world scenario.
162: Although this expectation is borne out, the surprise in these
163: scenarios is always the comparative difficulty of finding good
164: constraints in the event that the only known particle living in
165: the bulk is the graviton, since so little is directly known about
166: the graviton's properties. Some comparatively weak limits do
167: exist, coming from Parameterized Post-Newtonian tests of General
168: Relativity within the solar system \cite{WillBook,Will}, and from
169: its success in describing the energy loss of the binary pulsar,
170: 1913+16 \cite{Will,Taylor,Carlip}.
171: 
172: Much stronger limits on changes to the graviton dispersion
173: relation may also be obtained if they permit high-energy particles
174: to \v Cerenkov radiate gravitons. In this case bounds may be obtained
175: from our observation of very high energy particles in cosmic rays.
176: Assuming these cosmic rays to be protons -- as the best evidence
177: indicates \cite{protvsphot} -- the constraint $(c_p - c_g)/c_p <
178: 10^{-15}$ was obtained in this way \cite{MN}, strongly restricting the
179: case where the graviton velocity, $c_g$, is smaller than that of the
180: proton, $c_p$.
181: 
182: Our purpose in this paper is to provide a complementary constraint
183: for the case where protons cannot radiate gravitons, and so the
184: above bound does not apply. We obtain our constraint by computing
185: some of the Lorentz-violating effects which are implied for
186: fermions and photons by radiative corrections involving gravitons.
187: Because the bounds on such corrections for electrons and photons
188: are extremely good, we are able to infer reasonably strong bounds
189: on Lorentz violation in the graviton sector.
190: 
191: We find two kinds of effects, each of which depends on physics at
192: different scales. We find the one-loop graviton-induced
193: contributions to particle phase velocities to be highly suppressed
194: by the mass of the particle involved. This sensitivity is stronger
195: the higher the number of dimensions the graviton sees. The photon
196: and electron are particularly insensitive to this kind of
197: correction. Because they are so mass-dependent, the largest
198: contributions to light particles are often due to higher-loop
199: graphs within which a one-loop graviton-induced modification to
200: the propagation of a heavy particle (like a $W$ boson or top
201: quark) is inserted. This makes predictions sensitive to the
202: ultraviolet structure of the theory.
203: 
204: Graviton loops also change light-particle dispersion relations
205: by introducing contributions which involve higher powers of the
206: particle momentum. We compute these and find that they
207: can be generically less ultraviolet sensitive
208: than are changes to particle phase velocities. Interestingly, the
209: low-energy contributions of this form ({\it i.e.} the contributions
210: due to the massless 4D graviton) are already large enough to provide
211: interesting bounds.
212: 
213: We organize our presentation as follows. In the next section we
214: discuss in general terms how Lorentz-violating effects enter into
215: the low-energy 4D theory at energies well below the
216: compactification scale. This is the regime of interest for
217: experiments, and is {\it not} the regime in which one can consider
218: bulk particles to be moving on ballistic trajectories through the
219: extra dimensions. Then in section III we discuss the fermion and
220: photon self energies in general, to identify which features are
221: required in order to make comparison with experiments. Section IV
222: follows with the calculation of the graviton loop, and the
223: derivation of the induced Lorentz violation amongst visible
224: brane-based particles. Given the mass dependence of the results
225: of Section IV, Section V gives conservative estimates as to the
226: size which might be expected for photons, electrons and protons.
227: Section VI describes the bounds on Lorentz violation for these
228: particles, and computes the bound which may be inferred indirectly
229: on the strength of Lorentz violation in the gravity sector. Our
230: conclusions are summarized in section VI, and some of the
231: cumbersome intermediate results are gathered into two appendices.
232: 
233: \section{The 4d Effective Picture}
234: 
235: In the brane-world picture photons and electrons usually are
236: constrained to move on our four-dimensional brane, while gravitons are
237: free to explore the higher-dimensional bulk surrounding space. It is
238: intuitive in this kind of picture that brane and bulk particles might
239: propagate differently, since bulk particles might be free to take
240: `shortcuts' through the extra dimensions which are forbidden to
241: brane-bound particles \cite{Shortcuts}. This possibility has been
242: proposed to be, in some instances, a virtue in that it may provide a
243: novel way to address some cosmological problems \cite{DRV,CosmoShort}.
244: 
245: Although this kind of ballistic picture is appealingly intuitive,
246: it is not really appropriate for the low energies at which tests of
247: the dispersion relation actually take place. Experimental tests only involve
248: the one graviton which we see in 4 dimensions, and so only involve
249: the very lowest Kaluza-Klein (KK) state. By contrast, assigning the graviton
250: localized trajectories in the extra dimensions presupposes a localized
251: wave packet in these dimensions, which cannot be constructed purely
252: from the lowest-energy mode.
253: 
254: We briefly detour in this section to describe how extra-dimensional
255: Lorentz-violation appears in the low-energy effective Lagrangian
256: which describes the lowest KK mode. Although we start by using the
257: simplest example of a scalar field, the conclusions we draw will be
258: shown to be equally valid for higher spin fields.
259: 
260: \subsection{A Scalar Field}
261: Consider, therefore, a bulk scalar field, $\Phi(x,y)$, where $x^\mu$
262: labels the 4 dimensions parallel to the brane, and $y^m$ labels
263: the $n$ various transverse dimensions. The four-dimensional field
264: content is obtained by resolving $\Phi$ into a basis of modes
265: in the extra dimensions:
266: %
267: \eq
268: \Phi(x,y) = \sum_k \varphi_k(x) \, u_k(y),
269: \eeq
270: %
271: where the basis functions, $u_k(y)$, are eigenfunctions of the
272: appropriate kinetic/mass operator: $\Delta u_k = \omega_k \, u_k$.
273: 
274: The kinetic 4D action for the KK modes, $\varphi_k(x)$, is
275: obtained by inserting the mode expansion into the higher-dimensional
276: action and integrating the result over the
277: extra dimensions. Using an assumed form for the extra-dimensional
278: background metric:
279: %
280: \eq
281: \cG_{MN} = \pmatrix{ g_{\mu\nu}(x,y) & 0 \cr
282: 0 & h_{mn}(y) \cr}
283: \eeq
284: %
285: one finds in this way:
286: %
287: \eqa
288: S &=& -\, \frac12 \int d^{\,4}x \, d^{\,n}y \; \sqrt{-\cG} {\cG}^{MN}
289: \partial_M \Phi \partial_N \Phi \nonumber\\
290: &=& - \, \frac12 \int d^{\,4}x \; \sqrt{-G} \,
291: G^{\mu\nu}_{kl} \; \partial_\mu \varphi_k \partial_\nu \varphi_l
292: + \cdots
293: \eeqa
294: %
295: where $\cG = \det{\cG_{MN}}$, {\it etc.,} and the effective
296: four dimensional metric governing the kinetic terms is
297: %
298: \eq
299: \label{EffG}
300: \sqrt{-G}\, G^{\mu\nu}_{kl}(x) := \int d^{\,n} y\;
301: \sqrt{-gh}\,  g^{\mu\nu}(x,y) \; u^*_k(y) \,
302: u_l(y).
303: \eeq
304: 
305: The main point is that the metrics (plural since there is a
306: different metric for each choice of the indices $k,l$) defined by
307: eq.\ \pref{EffG} differ, in general, from the induced metric on
308: the brane, $\gamma_{\mu\nu}$, which appears in the kinetic term
309: for fields which are trapped on a brane. For instance, for a brane
310: defined by the surface $y=y_0$ the induced metric is simply
311: $\gamma_{\mu\nu} = g_{\mu\nu}(x,y_0)$. This is ultimately the
312: source of Lorentz-violating effects due to the bulk metric.
313: 
314: If we focus purely on the massless four-dimensional
315: mode, which we label by $k=0$, then we must integrate
316: out the other, more massive, KK modes. The kinetic term 
317: for the massless field,
318: $\varphi = \varphi_0(x)$, contains the metric
319: $G^{\mu\nu} = G^{\mu\nu}_{00}$ which may differ
320: from the induced metric on the brane, $\gamma^{\mu\nu}$.
321: In the rest of this paper we will focus on the implications
322: of this difference.
323: 
324: Two limits of the metric $G^{\mu\nu}$ bear highlighting.
325: First, in the absence of warping of the bulk metric
326: ({\it i.e.,} if $g^{\mu\nu}(x)$ is independent of $y$)
327: eq.\ \pref{EffG} becomes:
328: %
329: \eqa
330: \sqrt{-G}\, G^{\mu\nu}_{kl} &=& \sqrt{-g} \, g^{\mu\nu}(x)
331: \int d^{\,n}y \; \sqrt{h} \, u^*_k \, u_l \nn\\
332: &=& \delta_{kl} \; \sqrt{-g} \, g^{\mu\nu}(x), \eeqa
333: %
334: where the second line uses the orthonormality condition
335: of the basis modes, $u_k(y)$. In this case bulk and
336: brane modes see the same metric, for branes defined by
337: surfaces $y=y_0$.
338: 
339: Second, in the absence of a preferred frame in the
340: bulk metric (such as the AdS metric used by Randall
341: and Sundrum) the metrics $G_{\mu\nu}$ and $\gamma_{\mu\nu}$
342: must be conformal to one another ({\it i.e.,}
343: $G_{\mu\nu}(x) = f(x) \, \gamma_{\mu\nu}(x)$),
344: since Lorentz invariance implies they must both locally
345: be proportional to the Minkowski space metric $\eta_{\mu\nu}$.
346: 
347: \subsection{The 4D Graviton} A similar story gives the influence of
348: Lorentz-violating effects on the propagation of low-energy gravitons
349: within the effective theory below the compactification scale (or the
350: AdS curvature scale, in the case of RS-II-like models without
351: compactification \cite{RSII}), although the details are a bit more
352: complicated due to gauge invariance. Our purpose here is to identify
353: the leading contributions of Lorentz violations in the higher-energy
354: theory to graviton propagation in the low-energy 4D theory, and to show
355: that they always may be cast in terms of an appropriate shift of the
356: background metric.\footnote{We thank M. Pospelov for helpful
357: conversations on this point.}
358: 
359: We assume: ($i$) our interest is in energies very low compared to
360: the compactification scale, allowing a treatment in terms of the
361: low-energy 4D effective theory; ($ii$) there is only a single
362: massless spin-2 graviton mode in this effective theory, and
363: ($iii$) that the dominant effect of the higher-dimensional theory
364: is to break Lorentz invariance but not translation invariance or
365: rotational invariance (in the preferred frame). Under these
366: circumstances the 4D effective theory
367: involves an effective 4D metric field coupled to an order
368: parameter, $u_\mu$, which defines the preferred frame. The
369: assumptions of unbroken rotational invariance imply $u_\mu$ is
370: timelike, and we rescale it so that it is normalized, satisfying
371: $g^{\mu\nu} u_\mu u_\nu = -1$.
372: 
373: On grounds of general covariance, the 4D effective theory with
374: the fewest derivatives has the form:\footnote{Our metric is
375: `mostly plus', and we follow Weinberg's curvature conventions
376: \cite{GravCosm}.  Momentum 4-vectors are upper case ($P$),
377: the spatial vector is boldface (${\bf p}$), the magnitude of the spatial
378: piece is lowercase ($p$).}
379: %
380: \eq \label{d4DEH} {\cal L} = -\, {1 \over 2 \kappa_4^2} \;
381: \sqrt{-g} \Bigl[ R + a \; R^{\mu\nu} \, u_\mu u_\nu \Bigr] +
382: \dots \, ,\eeq
383: %
384: for some dimensionless constant $a$. Here the ellipses indicate
385: higher-derivative terms, and we have not written a cosmological
386: constant term, which we assume to be negligible. (We shed no light
387: in this paper on the vexing cosmological constant problem.) 
388: Here $\kappa_4^2 = 8 \pi \, G_4 = 1/M_4^2$, where
389: $G_4$ and $M_4$ are the usual 4D Newton's constant and
390: (rationalized) Planck mass: $M_4 \sim 2 \times 10^{18}$ GeV. In
391: order of magnitude we expect $a \sim \kappa^2_4 \Lambda^2$, where
392: $\Lambda$ is the scale associated with the Lorentz-violating
393: effects at higher energies. Clearly $a$ is naturally very small to
394: the extent that $\Lambda$ is much smaller than the 4D Planck
395: scale.
396: 
397: The main point now follows. The second term of eq.\ \pref{d4DEH}
398: may be completely absorbed by performing the following field
399: redefinition:
400: %
401: \eq \label{fieldredef} g_{\mu\nu} \to g_{\mu\nu} - \, {a \over 2}
402: \; \Bigl[ g_{\mu\nu} + 2 \, u_\mu u_\nu \Bigr] . \eeq
403: %
404: After having performed this redefinition (and a constant rescaling
405: of the metric), graviton fluctuations about flat space are
406: described performing the expansion $g_{\mu\nu}(x) = G_{\mu\nu} + 2
407: \kappa_4 \, h_{\mu\nu}(x)$, in eq.\ \pref{d4DEH}, using
408: %
409: \eq \label{metricform} G_{\mu\nu} = \eta_{\mu\nu} - \delta c_g^2
410: \, u_\mu u_\nu, \eeq
411: %
412: where $\eta_{\mu\nu} = \hbox{diag} (-1,1,1,1)$ is the usual
413: Minkowski metric, $u_\mu = \eta_{\mu\nu} u^\nu$, and $\delta
414: c_g^2$ is a small quantity which we shall see has the
415: interpretation as a change in the maximum propagation speed,
416: $c_g$, of the graviton. The field $h_{\mu\nu}$ here is the
417: canonically-normalized field describing graviton propagation.
418: 
419: Just as was the case for the scalar field, the leading effect of
420: higher-energy Lorentz violation is in this way seen to be a
421: modification of the background metric through which the graviton
422: propagates.
423: 
424: As a concrete illustration of this general argument, we can
425: compute the effective 4-D gravitational metric corresponding to
426: eq.\ (\ref{ads}), treating the Lorentz-violating term as a
427: perturbation.  The 4D effective gravitational action can be
428: obtained by expanding the 5D action $S$ to linear order in $\delta
429: g_{\mu\nu} = (-\mu/r^2+Q^2/r^4)u_\mu u_\nu$, and integrating over
430: the extra dimension.  If we write
431: %
432: \eq
433:     ds^2 \cong {r^2\over l^2}
434:     \left( g_{\mu\nu}(x) - {l^2\over r^2}\delta g_{\mu\nu}
435:     \right) dx^\mu dx^\nu +  {l^2\over r^2} dr^2
436: \eeq
437: %
438: (the correction $\Delta(r)$ in $g_{rr}$ can be neglected to leading order
439: in $\Delta$) then\footnote{For a warped metric of the form $ds^2 =
440: a(r)g_{\mu\nu}dx^\mu dx^\nu + b(r) dr^2$, the reduction of the
441: gravitational action from 5D to 4D is $S = -\frac12 M_5^3
442: (\int a\sqrt{b}\,dr)\sqrt{-g}R$, where $R$ is the Ricci scalar constructed
443: from the 4D part of the metric, $g_{\mu\nu}$.}
444: %
445: \eqa
446:     \delta  S &=& \frac12 M_5^3 \left(\int_{r_1}^{r_2} dr\,
447:     {r\over l}\cdot{l^2 \over r^2}\left[-{\mu\over r^2}+{Q^2\over r^4}
448:     \right]\right)
449:     \nonumber\\
450:     &&\times\sqrt{-g}\left(R^{\mu\nu}-\frac12 g^{\mu\nu}R\right) u_\mu u_\nu
451: \eeqa
452: %
453: The upper limit of integration corresponds to the position of one
454: brane, and the lower limit might be that of another brane, or else
455: the position of an event horizon where $\a(r_1)+\Delta(r_1)=0$, if
456: there is only a single brane.  Comparing to the preceding
457: discussion, we see that the graviton sees a metric of the form
458: (\ref{metricform}), with
459: %
460: \eq
461:     \delta c_g^2 \sim {\int_{r_1}^{r_2} dr\,
462:     {l \over r}\left[-{\mu\over r^2}+{Q^2\over r^4}
463:     \right] \over \int_{r_1}^{r_2} dr\, {r\over l} }
464: \eeq
465: %
466: Interestingly, the sign of $\delta c_g^2$ can be positive or negative,
467: depending on the relative sizes of black brane mass and charge.
468: 
469: \subsection{Physical Implications}
470: The physical significance of the metric, $G_{\mu\nu}$, appearing
471: in the kinetic term of a field within the effective theory, is most
472: easily seen by working within the geometrical-optics
473: approximation. Within this approximation, the propagating field is
474: written in the form $\varphi(x) = A(x) \, \exp[i P_\mu x^\mu]$,
475: with $A(x)$ assumed to be much more slowly-varying than is the
476: phase, $P_\mu x^\mu$. With this choice the field equation,
477: $(G^{\mu\nu} \nabla_\mu \nabla_\nu - m^2)\varphi = 0$ is
478: equivalent to the dispersion relation: $G^{\mu\nu} \, P_\mu P_\nu
479: + m^2 \approx 0$, or equivalently the normal vectors, $P_\mu$, of
480: the surfaces of constant phase are timelike (or null, if $m=0$)
481: vectors of the metric $G^{\mu\nu}$.
482: 
483: If the (four-dimensional) wavelength of the mode is much smaller
484: than the (four-dimensional) radius of curvature of the background
485: fields, then the motion of these wave packets is along the
486: geodesics of the 4D-metric $G_{\mu\nu}$. Clearly these
487: trajectories and dispersion relations generically differ for
488: fields which have different metrics in their kinetic terms.
489: 
490: The statement is slightly weaker for massless particles, since
491: the latter move along null geodesics of their respective
492: metrics. Consequently, their trajectories only differ if the two
493: metrics are not conformal to one another. In particular,
494: differences in the trajectories of massless particles are not
495: observable (in the geometric-optics limit) in the absence of a
496: preferred frame in the bulk or on the brane.
497: 
498: \section{Loops: General Considerations}
499: We now turn to the general implications for fermions and photons
500: of loop-generated Lorentz-violating effects. Motivated by the
501: considerations of the previous section, we imagine from here on
502: that all brane fields -- {\it i.e.,} all experimentally observed
503: elementary particles except for the graviton --  see only the
504: induced metric on the brane, which we take to be flat and Lorentz
505: invariant: $\gamma_{\mu\nu} = \eta_{\mu\nu} = \hbox{diag}
506: (-1,1,1,1)$. By contrast, the metric appearing within the kinetic
507: terms of any low-energy bulk fields -- which we take to be just
508: the graviton -- involves the Lorentz-violating metric of
509: eq.\ \pref{metricform}. This is the dominant low-energy source of
510: Lorentz violation in the effective theory, and it is the only type
511: of Lorentz violation whose implications we shall follow.
512: 
513: In general, loops involving virtual bulk states communicate the
514: news of Lorentz violation to the brane fields, and our task is to
515: compute the size of this effect. In this section we address
516: general issues which follow purely from the assumption that
517: $G^{\mu\nu} = \hbox{diag}(-1/c_g^2,1,1,1)$ encodes all
518: Lorentz-violating effects, and return in later sections to the
519: explicit calculation of these effects from graviton loops.
520: 
521: Since we know from direct bounds that Lorentz-violating bulk
522: effects are small (more about this later), we take $c_g = 1 +
523: \eps$ with $\eps \ll 1$. Because of the very strong constraints
524: already known for $c_g < 1$ \cite{MN} our primary interest in what
525: follows is in positive $\eps$. In view of the direct bounds
526: arising from solar-system and binary-pulsar tests of general
527: relativity we imagine $\eps \lsim 10^{-6}$\footnote{A weaker
528: bound, $\eps \lsim 10^{-3}$, is required if only terrestrial 
529: bounds, or the gravitational radiation rate of the binary pulsar are used. 
530: The stronger limit follows from angular momentum conservation 
531: for the Sun, as inferred by requiring the ecliptic and solar equatorial 
532: planes not to precess relative to one another throughout the
533: history of the solar system \cite{Nordtvedt,WillBook,Will}.}.
534: 
535: \subsection{Photon Propagation}
536: We identify in this section those parts of the graviton-induced
537: vacuum polarization which have implications for the dispersion
538: relation of transversely-polarized photons.
539: 
540: We first write the most general form for the vacuum polarization
541: which can be built from the tensors $\eta_{\mu\nu}$,
542: $G_{\mu\nu}=\eta_{\mu\nu} + (1-c_g^2) u_\mu u_\nu$ and the
543: momentum 4-vector, $P_\mu$, which is consistent with symmetry
544: ($\Pi^{\mu\nu} = \Pi^{\nu\mu}$) and transversality ($P_\mu \,
545: \Pi^{\mu\nu} = 0$). The most general form is:
546: %
547: \eqa \label{GenPi} \Pi^{\mu\nu} &=& A\; \left(\eta^{\mu\nu} -
548: {P^\mu P^\nu \over P^2} \right) + B \; \Bigl[ u^\mu u^\nu \\
549: && \quad \left. + {(P\cdot u)^2 \over P^4} \, P^\mu P^\nu  - \,
550: {(P \cdot u) \over P^2} \, (P^\mu u^\nu + P^\nu u^\mu) \right],
551: \nn \eeqa
552: %
553: where at this point $A$ and $B$ are arbitrary functions of the two
554: independent variables, $P^2$ and $P\cdot u$.
555: 
556: The dispersion relation is found by searching for the zero
557: eigenvalues of the inverse propagator, $\Delta^{\mu\nu} =
558: \Delta_0^{\mu\nu} + \Pi^{\mu\nu}$ where $\Delta_0^{\mu\nu} = -
559: (P^2 \, \eta^{\mu\nu} - P^\mu P^\nu)$. In particular, our interest
560: is only in those which are transverse (orthogonal to the pure
561: gauge directions). Working in the rest frame, $u^\mu = (1,0,0,0)$,
562: and taking the photon momentum to point in the $z$-direction ({\it
563: i.e.,} $P^\mu = (\omega,0,0,p)$), we therefore require that
564: $\Delta^{\mu\nu}=0$ in the directions $\mu,\nu = x,y$.
565: 
566: A simple calculation using eq.\ \pref{GenPi} shows that this
567: implies:
568: %
569: \eq \omega^2 - p^2 + A(\omega,p) = 0. \eeq
570: %
571: Since $A$ is perturbatively small, it suffices to write the
572: dispersion relation as $\omega = \omega_0 + \omega_1$, where
573: $\omega_0 = p$, and to evaluate $A$ with $\omega=\omega_0$. This
574: leads to the present section's main result:
575: %
576: \eq \label{phDispCond} \omega_1 = - \, {1 \over 2 \omega_0} \;
577: A(\omega_0,p). \eeq
578: 
579: As we shall find, $A$ admits an expansion at low energies in
580: powers of $P^2$ and $u\cdot P$, leading to the form
581: %
582: \eq A(\omega_0,p) = \alpha \, p^2 + \beta \, p^4 + \dots \, , \eeq
583: %
584: where rotational invariance precludes the appearance of odd powers
585: of $p$. Using this in eq.\ \pref{phDispCond} and comparing the
586: result with the general 4D photon dispersion relation
587: %
588: \eq \label{PhForm} \omega^2(p) = p^2 c_\gamma^2 + b_\gamma \,
589: p^{4} + O(p^6) \, . \eeq
590: %
591: we readily identify
592: %
593: \eq c_\gamma^2 = 1 - \alpha, \qquad b_\gamma = - \beta \, . \eeq
594: 
595: \subsection{Fermion Propagation}
596: We next ask how loop contributions to fermion self-energies can
597: modify fermion dispersion relations. We work within the rest frame
598: defined by $u^\mu$. Writing the inverse electron propagator as $S
599: = S_0 + \Sigma$, with $S_0 = - ( i \psl + m)$, we see that to
600: leading order in perturbation theory the zeroes of $S$ satisfy
601: $P^\mu = (E,\bp)$, where $E= E_0 + E_1$ with $E_0 = \sqrt{p^2 +
602: m^2_0}$ and:
603: %
604: \eq
605: \label{DispCond}
606: i \g^0 E_1 = - \Sigma(E_0,\bp).
607: \eeq
608: 
609: For small $\bp$, $\Sigma$ has the expansion
610: %
611: \eqa \label{GenSigma}
612: \Sigma(E_0,\bp) &=& \scA + \scB(i \vec\g
613: \cdot \bp) +
614: \scC \, p^2 \nn\\
615: && \qquad + \scD\, p^2 (i \vec\g \cdot \bp) + \scE \, p^4 +
616: \cdots \eeqa
617: %
618: and so eq.\ \pref{DispCond} implies a dispersion relation
619: of the form
620: %
621: \eq \label{disprel} E_f^2 = m^2_f + p^2\, c_f^2 + b_f \, p^4 +
622: \cdots \eeq
623: %
624: with
625: %
626: \eqa
627: m_f &=& m_0 - \scA + \cdots \\
628: c^2_f &=& 1 - 2(\scB + m_0 \, \scC) + \cdots \\
629: b_f &=& - 2 (\scD + m_0 \, \scE) + \cdots \eeqa
630: %
631: In these last three equations the subscript $f$ denotes the
632: fermion species, and the ellipses indicate higher-order
633: contributions.
634: 
635: Since it is the quantities $c_f^2$ and $b_f$ which we wish to
636: compare with experiments, the implications of graviton loops may
637: be obtained by computing the coefficients $\scB$ through $\scE$.
638: 
639: \section{Loops: Graviton Calculations} In this section we compute the
640: one-loop self energy which is obtained when a fermion or photon
641: emits and reabsorbs a virtual graviton, as in figs.\ 1 and 2. We
642: present our results in three steps. First, since the integrals
643: involved strongly diverge in the ultraviolet, we make some general
644: remarks about the correct way to treat these divergences before
645: describing the calculations themselves. We then evaluate the
646: graviton loop in two steps, motivated by the picture that Lorentz
647: violation is arising from field configurations within the
648: extra-dimensional bulk. First we consider loops involving only the
649: lowest KK mode: the massless 4D graviton. These loops have the
650: virtue of only involving known particles and couplings, and so the
651: results we obtain are comparatively robust. They describe the
652: graviton contributions within the effective 4D theory, well below
653: the compactification scale, $M_c \sim 1/r$, where $r$ is a measure
654: of the linear size of the extra dimensions. In the case of a
655: single noncompact extra dimension, the effective theory is good
656: below the bulk curvature scale $\sqrt{-\Lambda_5/M_5^3}$, where
657: $\Lambda_5$ is the bulk cosmological constant and $M_5$ is the 5-D
658: gravity scale.
659: 
660: Next, we compute the contributions of gravitons in the effective
661: theory between $M_c$ and the scale $M_l > M_c$ associated with the
662: extra-dimensional Lorentz violating physics. Since the theory is
663: extra-dimensional in this energy range, this involves calculating
664: the loop contributions of higher KK graviton modes. In order to do
665: this we make several simplifying assumptions about the nature of
666: the extra-dimensional Lorentz violation, which we believe suffices
667: for the purposes of establishing the order of magnitude of the
668: extra-dimensional result.
669: 
670: \subsection{Ultraviolet Divergences}
671: In $d$ spacetime dimensions the gravitational coupling has
672: dimension $\kappa_d \sim M_d^{1-d/2}$, where $M_d$ is the
673: $d$-dimensional Planck mass. On dimensional grounds we therefore
674: expect the most divergent contribution to one-loop brane-particle
675: dispersion relations to be
676: %
677: \eq c^2 - 1 \sim \kappa_d^2 \, \Lambda^{d-2}, \qquad b \sim
678: \kappa_d^2 \, \Lambda^{d-4},  \eeq
679: %
680: where $\Lambda$ is the ultraviolet cutoff scale.
681: 
682: As is usual within an effective theory, this indicates that the
683: result is most sensitive to the most energetic degrees of freedom
684: in the problem, suggesting that calculations within the full
685: theory would produce results that are of order $c^2 -1 \sim
686: \kappa_d^2 \, M^{d-2}$ and $b \sim \kappa_d^2 \, M^{d-4}$, where
687: $M$ might be the mass of a heavy particle which was integrated out
688: to produce the low-energy effective theory.
689: 
690: As we shall see, the above mass-dependence is roughly right,
691: although some care is required due to the appearance of power-law
692: divergences \cite{BL}. Care is required because, although the
693: renormalization group ensures that the coefficient of large
694: logarithms like $\log(M)$ in observable quantities like $c^2 -1$
695: or $b$ may be read off from the coefficients of log divergences
696: (like $\log(\Lambda)$) within the effective theory, the same is
697: not true for higher (power-law) divergences. As a result in this
698: section we ignore all power divergences, and compute only the
699: log-divergent parts of the results. If the results do not have log
700: divergences (as will be the case with an odd number of extra
701: dimensions), then we compute only the finite parts of the loops.
702: 
703: Neglecting the power divergences minimizes the Lorentz symmetry
704: violation which is seen by particles on the brane, and so leads to
705: conservative conclusions. It corresponds to considering the theory
706: in which brane-bound particles respect Lorentz invariance at the
707: energy scale where the theory becomes 4 dimensional. We return in
708: Section V to the issue of contributions which are proportional to
709: positive powers of large mass scales, $M$, by considering higher
710: loops which explicitly involve more massive virtual particles. We
711: shall there see that naive power-counting estimates do correctly
712: reproduce the $M$ dependence of the results, but miss important
713: dimensionless loop factors.
714: 
715: In practice the finite and log-divergent terms are most easily
716: obtained within dimensional regularization, within which power-law
717: divergences do not arise. We have computed our results both using
718: dimensional regularization and using an explicit ultraviolet
719: cutoff, however, and have verified that the answers obtained are
720: the same in both cases.
721: 
722: \subsection{Four-dimensional Graviton}
723: The Feynman rules for fermions and gravitons in the absence of
724: Lorentz-violating effects are standard, and are obtained by
725: linearizing the Dirac-Einstein-Hilbert action in curved space,
726: %
727: \eq \label{DEHAction} {\cal L} = - \sqrt{-g} \, \left[ {1 \over 2
728: \kappa_4^2} \; R + \overline\psi (\dsl + m) \psi + {1 \over 4} \;
729: F_{\mu\nu} \, F^{\mu\nu} \right], \eeq
730: %
731: about flat space: $g_{\mu\nu} = \eta_{\mu\nu} + 2 \kappa_4 \,
732: h_{\mu\nu}$. As before $\kappa_4$ denotes the rationalized 
733: Planck mass in 4D: $M_4 \sim 10^{18}$ GeV. $h_{\mu\nu}$
734: represents the canonically-normalized graviton field. A recent
735: statement of the resulting Feynman rules can be found in
736: references \cite{GRW,HLZ}.
737: 
738: As we have argued in previous sections, we know that the only
739: Lorentz-violating modification to these rules consists in
740: replacing the Minkowski metric, $\eta_{\mu\nu}$, with the
741: nonstandard metric, $G_{\mu\nu}$, when linearizing the first term
742: in eq.~\pref{DEHAction} to obtain the graviton propagator. In de
743: Donder gauge this gives:
744: %
745: \eq \label{GravProp} G_{\mu\nu:\alpha\beta}(Q) =
746: {P_{\mu\nu:\alpha\beta} \over G^{\lambda\rho}\, Q_\lambda Q_\rho -
747: i\veps} \, , \eeq
748: %
749: where
750: %
751: \eq P_{\mu\nu:\alpha\beta} = \half \left( G_{\mu\alpha}\,
752: G_{\nu\beta} + G_{\mu\beta} \, G_{\nu\alpha} - G_{\mu\nu} \,
753: G_{\alpha\beta} \right) ,\eeq
754: %
755: and $\veps$ -- not to be confused with $\eps = c_g-1$ -- is the
756: infinitesimal which ensures the propagator satisfies Feynman
757: boundary conditions. The fermion propagators and vertices arise on
758: the brane, and so use only the usual Minkowski metric.
759: 
760: \subsubsection{The Photon Vacuum Polarization}
761: 
762: 
763: 
764: \medskip
765: \centerline{\epsfxsize=2.5in\epsfbox{photon.eps}}
766: {\small
767: Figure 1. Graviton contribution to photon vacuum polarization.}
768: \medskip
769: 
770: 
771: After Wick rotation, the one-loop Feynman graph in which a virtual
772: graviton is emitted and reabsorbed by the photon (fig.\ 1) leads
773: to the following expression for the photon vacuum polarization:
774: %
775: \eq \Pi^{\mu\nu} = \left( {\kappa_4 \over 2} \right)^2 \int {d^4 Q
776: \over (2\pi)^4} \; {N^{\mu\nu} \over D} , \eeq
777: %
778: where $D = (P-Q)^2 \; G^{\alpha\beta}Q_\alpha Q_\beta$ and
779: %
780: \eq N^{\mu\nu} = V^{\mu\lambda:\alpha\beta}(P,P-Q) \;
781: {V_\lambda}^{\nu: \sigma\rho}(P-Q,P) \; P_{\alpha\beta:\sigma\rho}
782: \, . \eeq
783: %
784: Appendix A gives explicit expressions for the vertex functions,
785: $V^{\alpha\beta:\mu\nu}(P,Q)$. In these expressions all dependence
786: on the metric $G^{\mu\nu}$ is explicit, and the brane metric,
787: $\eta_{\mu\nu}$, is to be used to perform any implicit index
788: contractions, such as in $(P-Q)^2$.
789: 
790: Evaluating this expression (we used the programs FORM and
791: MATHEMATICA to perform the tensor contractions), Taylor expanding
792: in powers of $\eps = c_g - 1$ and performing the momentum integral
793: gives the following expression for the coefficient function, $A$,
794: of eq.\ \pref{GenPi}:
795: %
796: \eq A(\omega_0, p) = {304 \over 15} \; \lambda_4 \, \eps^2 \, p^4
797: + O(\eps^3), \eeq
798: %
799: where $\lambda_4 = (\kappa_4/8\pi)^2 \; {\cal L}$. Here ${\cal L}
800: = \log(\Lambda^2/\mu^2) = 2/(4-n)$, where the first equality is
801: evaluated using an ultraviolet cutoff, $\Lambda$, and the second
802: regularizes by continuing the spacetime dimension, $n$, away from
803: 4. $\mu$ is an arbitrary scale. We ignore the finite part of the
804: integral relative to its log-divergent part.
805: 
806: Comparison with the general expressions provided earlier gives the
807: following dispersion coefficients:
808: %
809: \eq \label{phresult}
810: c_\gamma^2 - 1 = 0, \qquad \qquad b_\gamma =
811: -\, {304\over 15} \, \lambda_4 \, \eps^2. \eeq
812: 
813: 
814: \subsubsection{The Fermion Self-Energy}
815: 
816: 
817: \medskip
818: \centerline{\epsfxsize=2.5in\epsfbox{fermion.eps}}
819: {\small
820: Figure 2. Graviton contribution to fermion self-energy.}
821: \medskip
822: 
823: 
824: We proceed in a similar way for the changes to the fermion self
825: energy. Evaluating the one-loop Feynman graph using the graviton
826: propagator of eq.\ \pref{GravProp}, and Wick-rotating to Euclidean
827: momenta, leads to the following expression for the fermion self
828: energy:
829: %
830: \eq \Sigma_4 = - \left ( {\kappa_4\over 2} \right)^2 \int {d^4 Q
831: \over (2\pi)^4} \; {N_1 \over D}, \eeq
832: %
833: where
834: %
835: \eqa \label{None}
836: N_1 &=& \frac12 \, G_{\mu\nu} \, \gamma^\mu [-i
837: (\psl - \qsl) + m] \gamma^\nu \nn\\
838: &&\qquad\qquad \times \; G_{\alpha\beta}(2 P - Q)^\alpha
839: (2P - Q)^\beta, \nn\\
840: D &=& G^{\lambda\rho} P_\lambda P_\rho \; [(P-Q)^2 + m^2]. \eeqa
841: %
842: Again all dependence on the metric $G^{\mu\nu}$ is explicit, and
843: the brane metric, $\eta_{\mu\nu}$, is used to perform the implicit
844: index contractions in $(P-Q)^2$ and $\psl$. To derive this we used
845: the simple vertex function $(k_1+k_2)_\mu \gamma_\nu +
846: (k_1+k_2)_\nu \gamma_\mu$ of ref.\ \cite{GRW} for the
847: fermion-fermion-graviton coupling rather than the more complicated
848: one of \cite{HLZ}, which includes an extra term $(i\ksl_1+i\ksl_2
849: +2m)\eta_{\mu\nu}$. The neglect of this extra term is justified --
850: even within loop graphs -- because it vanishes if the fermion
851: field equations are used. This ensures that it is an irrelevant
852: operator, in the sense that it can be removed by performing a
853: field redefinition of the fermion of the form $\delta \psi \propto
854: {h^\mu}_\mu \; \psi$.
855: 
856: We have evaluated this integral to the lowest two orders in the
857: small parameter $\eps = c_g - 1$, using the programs
858: FORM/MATHEMATICA to
859: keep track of the various 4-vectors in the problem. We
860: find the following results for the logarithmically-divergent part
861: of the coefficients $\scA$ through $\scE$ of eq.\ \pref{GenSigma}:
862: %
863: \eqa \scA_4 &=&  m^3 \lambda_4 \; \left(4 + 13 \eps + {33\over 2}
864: \, \eps^2 + ...\right) ,\nn\\
865:  \scB_4 &=&  m^2 \lambda_4 \; \left(4 \eps + 10 \eps^2 + ...\right) ,\nn\\
866:  \scC_4 &=&  {m \lambda_4\over 3}  \; \left(16 \eps + 35 \eps^2 + ...\right) ,\nn\\
867:  \scD_4 &=&   \lambda_4 \; \left(6 \eps^2 + ...\right) ,\nn\\
868:  \scE_4 &=& 0 + ... \; . \eeqa
869: %
870: where as before $\lambda_4 = (\kappa_4/8\pi)^2 \; {\cal L}$, with
871: ${\cal L} = \log(\Lambda^2/\mu^2) = 2/(4-n)$.
872: 
873: These imply the following contributions to the dispersion relation
874: of eq.\ \pref{disprel}:
875: %
876: \eqa c_f^2 -1 &=& -{2 m_f^2 \lambda_4 \over 3} \; \left[ 28 \eps +
877: 65 \eps^2 + ... \right], \nn\\
878: b_f &=& - \lambda_4 \left[ 12 \eps^2 + ... \right] \eeqa
879: %
880: For the case of interest, $\eps > 0$, we see that $c_f^2 < 1$
881: corresponding to fermions propagating {\it slower} than light. It
882: follows that (for a given momentum, $p$) the fermion energy is
883: depressed compared to the photon energy for small $p$.
884: 
885: Notice that the coefficients $b_f$ and $b_\gamma$ both first arise
886: at $O(\eps^2)$, and so their sign (which is negative for both)
887: does not depend on the sign of $\eps$. Furthermore, these terms
888: satisfy $b_f > b_\gamma$. These terms therefore act to raise the
889: fermion energy relative to the photon, and so act in the opposite
890: direction of the effect of $c_f < c_\gamma$ (when $\eps
891: > 0$).
892: 
893: \subsection{Higher-dimensional Gravitons}
894: %
895: We may now estimate the contributions of the higher KK graviton
896: modes to the propagation of brane-based fermions. Rather than
897: doing so by performing a sum over a tower of 4-D KK states, we
898: proceed by directly performing the loop graph using
899: higher-dimensional gravitons.
900: 
901: The first step is to specify what the higher-dimensional metric is
902: about which the graviton fluctuation is to be considered. In
903: principle this should be the metric which describes the
904: gravitational field of the object or objects in the bulk, whose
905: presence gives rise to the preferred frame which violates Lorentz
906: invariance. Since the explicit form for such metrics is rarely
907: known, we proceed by a more approximate route.
908: 
909: Our approximation is based on the observation that
910: higher-dimensional graviton loops are ultraviolet sensitive,
911: with their dominant contributions arising from the circulation of
912: very short-wavelength modes. So long as the wavelength of these
913: modes is much shorter than the radii of curvature of the
914: background Lorentz-violating metric, it should be sufficient to
915: replace the background metric by one which is approximately flat,
916: but Lorentz-violating. In particular, this should be sufficiently
917: accurate for our purposes of estimating the order-of-magnitude of
918: the resulting loop-generated contributions to fermion propagation
919: on the brane.
920: 
921: Accordingly we imagine the higher-dimensional graviton to
922: propagate about a flat metric in which there is a preferred frame
923: defined by an approximately constant $d$-vector, $u^\mu$. Such a
924: metric is again described by eq.\ \pref{metricform}, although with
925: $G_{\mu\nu}$ now being $d$-dimensional.
926: 
927: Linearizing the extra-dimensional Einstein-Hilbert action about
928: this metric -- $g_{ab} = G_{ab} + 2 \kappa_d \, h_{ab}$ -- we find
929: the following $d$-dimensional graviton propagator in de Donder
930: gauge:
931: %
932: \eq \label{GravPropd} G_{\mu\nu:\alpha\beta} = {\half \left(
933: G_{\mu\alpha}\, G_{\nu\beta} + G_{\mu\beta} \,
934: G_{\nu\alpha}\right) - {1\over d-2} \; G_{\mu\nu} \,
935: G_{\alpha\beta} \over G^{\lambda\rho} P_\lambda P_\rho -i\veps} \,
936: . \eeq
937: %
938: Again $\veps$ in this expression is the infinitesimal which
939: enforces Feynman boundary conditions. The quantity $\kappa_d$ is
940: related to the $d$-dimensional Newton's constant and Planck mass
941: by $\kappa_d^2 = 8 \pi G_d = (1/M_d)^{d-2}$. We take the fermions
942: and photons to move on a 3-brane, for which the induced metric is
943: the usual Minkowski metric.
944: 
945: \subsubsection{The Photon Vacuum Polarization}
946: Computing the photon vacuum polarization with this propagator
947: gives no correction to the photon dispersion relation at low
948: energies, if the number of dimensions exceeds the usual four.
949: This can be understood purely in terms of dimensional analysis,
950: and the fact that we are computing only the finite or log-divergent
951: contributions. The one-loop contributions to the photon
952: self-energy have two graviton vertices, so they are proportional
953: to $\kappa^2_d = M_d^{-(d-2)}$. Since the log-divergent and finite parts do not
954: involve powers of the cutoff $\Lambda$, the only quantity which
955: can be used to make a dimensionally correct answer is the photon
956: momentum, $p$. Thus the result is proportional to $\kappa_d^2
957: \, p^{d}$.  This gives a $p^4$ correction to the dispersion relation,
958: but only if $d=4$. For $d>4$ the correction always gives only a higher
959: than quartic power of $p$. (Section V discusses the physics of the
960: power-divergent contributions to $c_\gamma^2 - 1$ and $b_\gamma$.)
961: 
962: \subsubsection{The Fermion Self-Energy}
963: Evaluating the one-loop fermion self-energy graph using the
964: $d$-dimensional graviton propagator of eq.\ \pref{GravPropd}, and
965: Wick-rotating to Euclidean momenta, leads to the following
966: expression for the fermion self energy:
967: %
968: \eq \Sigma_d = - \left ( {\kappa_d\over 2} \right)^2 \int {d^d Q
969: \over (2\pi)^d} \; \left( {N_1 + N_2 \over D} \right), \eeq
970: %
971: where $N_1$ is as given by eq.\ \pref{None} and
972: %
973: \eqa N_2 &=& C_d \, \Bigl\{ [i(\psl {-} \qsl) {+} m ]
974: G_{\mu\nu} \, (2 P {-} Q)^\mu (2 P {-} Q)^\nu \nn\\
975: && \; \, -2i G_{\alpha\beta} \gamma^\alpha (2 P {-} Q)^\beta
976: G_{\lambda\rho} (P {-} Q)^\lambda (2P {-} Q)^\rho \Bigr\},\eeqa
977: %
978: with $C_d =  (d-4)/[2(d-2)]$.
979: 
980: Evaluating the integral in powers of $\eps = c_g - 1$ gives an
981: ultraviolet divergent result. We have evaluated the result using
982: both dimensional regularization and an explicit cutoff, and have
983: verified that the coefficients of the logarithmically-divergent
984: and pole terms are the same in both cases. For odd dimensions the
985: integrals are finite in dimensional regularization, and we have
986: verified that the result agrees with the finite part when
987: evaluated with an ultraviolet cutoff.
988: 
989: We are led in this way to the following expressions for the
990: quantity $c_f^2 - 1$ for dimensions $d=5$ through 10. (We give the
991: quantities $\scA$ through $\scE$ in Appendix B.):
992: %
993: \eqa c_f^2 {-}1 &=&  \!\phm m_f^3 \, \lambda_5  \: \left[ \frac{110}{9} \,
994: \eps + \frac{239}{9} \, \eps^2 + ... \right],  \quad (d=5)\nn\\
995: &=&\!  \phm m_f^4 \, \lambda_6  \: \left[ \frac{\: 48\: }{5}
996: \, \eps + \frac{\; 99 \;}{5} \, \eps^2 + ... \right],  \quad (d=6)\nn\\
997: &=&\! -\, m_f^5 \, \lambda_7  \: \left[ \frac{616}{75}
998: \, \eps + \frac{244}{15} \, \eps^2 + ... \right],   \quad (d=7) \nn\\
999: &=&\! -\, m_f^6 \, \lambda_8  \: \left[ \frac{464}{63}
1000: \, \eps + \frac{890}{63} \, \eps^2 + ... \right],  \quad (d=8)
1001: \nn\\
1002: &=&\!  \phm m_f^7 \, \lambda_9  \: \left[ \frac{333}{49}
1003: \, \eps + \frac{\!1245\!}{98} \, \eps^2 \!+ ... \right],  \quad (d=9) \nn\\
1004: &=&\! \phm m_f^8 \, \lambda_{10\!} \left[ \frac{115}{18} \, \eps +
1005: \frac{421}{36} \, \eps^2 + ... \right], \quad (d=10). \! \eeqa
1006: %
1007: Here the quantities $\lambda_d$ are defined in terms of the
1008: couplings $\kappa_d$ by:
1009: %
1010: \eqa \lambda_5 = {\kappa_5^2\over (8 \pi)^2}, \phl \quad \;\;
1011: \lambda_6 &=& {2\kappa_6^2\over (8 \pi)^3} \; {\cal L}, \quad \;\;
1012: \lambda_7 \,= {\kappa_7^2 \over 6(4\pi)^3}, \\
1013: \lambda_8 = {2\kappa_8^2 \over (8 \pi)^4} \; {\cal L}, \quad \;\;
1014: \lambda_9 &=& {\kappa_9^2\over 15 (4 \pi)^4}, \quad \;\;
1015: \lambda_{10\!\!} = {4\kappa_{10}^2 \over 3(8 \pi)^5} \; {\cal L}.
1016: \nn \eeqa
1017: %
1018: As before ${\cal L} = \log(\Lambda^2/\mu^2)$ when a cutoff is
1019: used, or ${\cal L} = 2/(d-n)$ in dimensional continuation of $n$
1020: away from $n=d$.
1021: 
1022: Notice that the corrections to $c_f^2$ which are implied in this
1023: way are not universal in size for all fermions, being suppressed
1024: by powers of $m_f$ for lighter fermions.
1025: 
1026: The corresponding higher-order dispersion coefficient, $b_f$, is
1027: similarly:
1028: %
1029: \eqa b_f &=&   \phm m_f \, \lambda_5 \,\left[ \frac{26}{3} \, \eps^2 +
1030: ... \right]  , \qquad (d=5) \nn\\
1031: &=&   \phm m_f^2 \, \lambda_6 \,\left[ \frac{36}{5} \, \eps^2 + ...
1032: \right] , \qquad (d=6) \nn\\
1033: &=& -\, m_f^3 \, \lambda_7 \,\left[ \frac{32}{5} \, \eps^2 + ... \right],
1034: \qquad (d=7) \nn\\
1035: &=& -\, m_f^4 \, \lambda_8 \,\left[ \frac{\! 124 \!}{21} \, \eps^2 \!+ ...
1036: \right] , \qquad (d=8) \nn\\
1037: &=&  \phm m_f^5 \, \lambda_9 \,\left[ \frac{39}{7} \, \eps^2 + ...
1038: \right] , \qquad (d=9) \nn\\
1039: \label{bfeqs}
1040: &=&  \phm m_f^6 \, \lambda_{10\!\!} \left[ \frac{16}{3} \, \eps^2 + ...
1041: \right] , \qquad (d=10) . \eeqa
1042: %
1043: Just as we saw for $d=4$, $b_f$ always arises at second order in
1044: $\eps$, and so its sign is completely determined in our
1045: calculation. As we shall see, the best bound on this coefficient
1046: arises when $b_f > b_\gamma \approx 0$.
1047: 
1048: The above expressions were derived specifically with the scenario
1049: of large, flat extra dimensions in mind.  However we can also
1050: interpret at least the 5D result in terms of a warped extra
1051: dimension.  This has more interesting consequences than the flat
1052: case, where the quantum gravity scale would have to exceed
1053: $M_5\gsim 10^8$ GeV in order to  comply with sub-millimeter tests
1054: of gravity \cite{adelberger}.  In the warped case, even if
1055: $M_5\sim M_p$, the KK gravitons couple to the  TeV brane with TeV
1056: strength, and they have a mass gap of order TeV.   In this model,
1057: we live on a ``TeV brane'' at $r=r_1$, displaced from the ``Planck
1058: brane'' at $r=r_2$ such that ${r_1\over r_2}\sim 10^{-16}$ in
1059: accordance with solving the hierarchy problem.  As far as the
1060: contributions from the ultraviolet graviton loops are concerned,
1061: this looks like quantum gravity with a scale of $M_5\sim$ TeV. The
1062: TeV mass gap protects low energy gravity from any observable
1063: distortion, but not so the loop effects from momenta $p\gg$ TeV.
1064: 
1065: \section{Real-World Complications}
1066: %
1067: A further step is required before these results can be compared
1068: with the experimental limits on the dispersion relations of real
1069: particles. This step involves identifying which low-energy
1070: particles produce the largest contributions to any given
1071: dispersion relation.\footnote{We thank M.~Pospelov for making many
1072: very useful suggestions for this and the next sections.}
1073: 
1074: There are two reasons why this additional step is required. First,
1075: we would like to apply the above calculations to protons, for
1076: which the experimental limits are the strongest. Unfortunately,
1077: our calculations treat all fermions as elementary, and so can only
1078: apply directly to the proton for scales below roughly $m_p \sim 1$
1079: GeV. (Notice these low energies may nonetheless be described by an
1080: extra-dimensional effective theory within ADD-type scenarios.) For
1081: higher energies, we must apply our calculations to the constituent
1082: quarks and gluons, and infer from these how the proton dispersion
1083: is affected. Unfortunately, the resulting strong-interaction
1084: uncertainties prohibit us from following in more detail the $O(1)$
1085: factors and signs of the results, limiting us to an
1086: order-of-magnitude analysis for the proton.
1087: 
1088: The second reason for being careful in applying our results is the
1089: strong dependence which they have on the relevant fermion masses. In
1090: particular, it may be that larger contributions are obtained for light
1091: particles (like electrons and photons) by embedding the
1092: graviton-induced Lorentz-violating contributions of heavier particles
1093: (like top quarks) within additional loops. This is how we will recover
1094: the larger contributions that might have naively been obtained by
1095: keeping the power divergences of the graviton loop graphs.
1096: 
1097: \subsection{Photons}
1098: %
1099: Our result for the photon dispersion relation is particularly
1100: simple, with $c_\gamma = 1$ for all $d$, and $b_\gamma \ne 0$ only
1101: for $d=4$. This simplicity is largely due to our concentrating on
1102: the finite and log-divergent contributions, however, which
1103: suggests that larger contributions may be found at the
1104: (comparatively cheap) expense of introducing additional loops.
1105: 
1106: For instance, at two loops the photon vacuum polarization acquires
1107: contributions by inserting a Lorentz-violating graviton
1108: self-energy within a charged-fermion loop (fig.\ 3). Assuming all
1109: charged fermions to be brane-bound, and so 4-dimensional, we
1110: estimate the following results for the photon: $c_\gamma^2 -1 \sim
1111: \left( {\alpha \over 4 \pi} \right) \, (c^2_f - 1)$, giving
1112: %
1113: \eqa \label{phest}
1114: c_\gamma^2 -1 &\sim& \left( {\alpha \over 4 \pi}
1115: \right) \left( {\eps \over (4 \pi)^{[d/2]}} \right) \; \left( {m_f
1116: \over
1117: M_d} \right)^{d-2} \\
1118: &\sim& \left\{ \matrix{ 1\times 10^{-10} \, \left( {\eps \times
1119: 10^{3}} \right) \, \left( { \hbox{TeV} / M_4} \right)^2 & (d=4)\cr
1120: 2 \times 10^{-11} \, \left( {\eps \times 10^{3}} \right) \, \left(
1121: { \hbox{TeV} / M_5} \right)^3 & (d=5) \cr
1122: 3 \times 10^{-13} \, \left( {\eps \times 10^{3}} \right) \, \left(
1123: { \hbox{TeV} / M_6} \right)^4 & (d=6) \cr} \right.  \, ,\nn \eeqa
1124: %
1125: where $[d/2]$ denotes the integer part of $d/2$ and we use the
1126: heaviest known elementary particle, the top quark ($m_t = 175$
1127: GeV), for numerical purposes. Although this is still negligibly
1128: small for $d=4$, we shall see that the $d>4$ result can be large
1129: enough to provide new constraints on $\eps$ if $M_d$ is not too
1130: far above the TeV range.
1131: 
1132: A similar contribution arises to $b_\gamma$, although because of
1133: the weaker dependence on $m_f$ the price of a loop factor,
1134: $\alpha/4\pi$, is not worthwhile when $d=4$, where the direct
1135: one-loop result of eq.~\pref{phresult} is larger. We expect,
1136: then
1137: %
1138: \eqa \label{phestbd4} b_\gamma M_4^2 &=& - \, {304\over 15} \left(
1139: {\eps \over 8 \pi} \right)^2 \qquad (d = 4)\\
1140: &=& -3 \times 10^{-8} \, \left( {\eps \times 10^{3}}
1141: \right)^2\eeqa
1142: %
1143: in four dimensions, where the numerical result uses the
1144: conservative estimate $\log(\Lambda^2/\mu^2) \sim 1$.
1145: 
1146: For $d>4$ we estimate the two-loop result by $b_\gamma \sim \left(
1147: {\alpha \over 4 \pi} \right) \, b_f$, and so find:
1148: %
1149: \eqa \label{phestb} b_\gamma M_d^2 &\sim& \left( {\alpha \over 4
1150: \pi} \right) \left( {\eps^2 \over (4 \pi)^{[d/2]} }
1151: \right) \; \left( {m_f \over M_d} \right)^{d-4} \qquad (d>4)\\
1152: &\sim& \left\{ \matrix{ 7 \times 10^{-13} \, \left( {\eps \times
1153: 10^{3}} \right)^2 \, \left( { \hbox{TeV} / M_5} \right) & (d=5)
1154: \cr 1 \times 10^{-14} \, \left( {\eps \times 10^{3}} \right)^2 \,
1155: \left( { \hbox{TeV} / M_6} \right)^2 & (d=6) \cr} \right.  \, ,\nn
1156: \eeqa
1157: %
1158: with $m_t$ being used for the fermion mass. The sign of $b_\gamma$
1159: and $c_\gamma^2 -1$ are not determined by these estimates
1160: (although they are certainly calculable within a more careful
1161: evaluation of Fig.~(3).
1162: 
1163: \medskip
1164: \centerline{\epsfxsize=2.5in\epsfbox{vacpol.eps}}
1165: {\small
1166: Figure 3. Two loop contribution to photon vacuum polarization from
1167: top quark and graviton.}
1168: 
1169: \subsection{Electrons}
1170: The next-cleanest application of the above analysis is to
1171: electrons, since so far as we know these are elementary and so our
1172: earlier results may be directly applied. Again taking the
1173: conservative estimate $\log(\Lambda^2/\mu^2) \sim 1$, we find that
1174: the direct graviton-loop contribution to the electron dispersion
1175: relation is
1176: %
1177: \eqa \label{eldir} c_e^2 - 1 &\sim& { \eps \over (4 \pi)^{[d/2]}}
1178: \; \left( {m_e
1179: \over M_d} \right)^{d-2}, \nn\\
1180: b_e &\sim& {\eps^2 \over (4 \pi)^{[d/2]} \, M_d^2} \; \left( {m_e
1181: \over M_d} \right)^{d-4}, \eeqa
1182: %
1183: where $d$ counts the number of spacetime dimensions seen by the
1184: graviton within the effective theory at scales $\mu < \Lambda \sim
1185: 1$ TeV.
1186: 
1187: This should be compared with the result of inserting more massive
1188: particles into higher loops. For instance, a loop with a W boson and
1189: neutrino, with the graviton coupling to the W (fig.\ 4)
1190: gives the alternative contribution
1191: %
1192: \eqa \label{elest} c_e^2 -1 &\sim& \left( {\alpha_w \over 4 \pi}
1193: \right) \left( {\eps \over (4 \pi)^{[d/2]}} \right) \; \left(
1194: {m_w \over M_d} \right)^{d-2} \\
1195: &\sim& \left\{ \matrix{ 1\times 10^{-10} \, \left( {\eps \times
1196: 10^{3}} \right) \, \left( { \hbox{TeV} / M_4} \right)^2 & (d=4)\cr
1197: 3 \times 10^{-12} \, \left( {\eps \times 10^{3}} \right) \, \left(
1198: { \hbox{TeV} / M_5} \right)^3 & (d=5) \cr
1199: 6 \times 10^{-14} \, \left( {\eps \times 10^{3}} \right) \, \left(
1200: { \hbox{TeV} / M_6} \right)^4 & (d=6) \cr} \right.  \, ,\nn \eeqa
1201: %
1202: using $m_w = 80$ GeV, and $\log(\Lambda^2/mu^2) = 1$.
1203: This last contribution (for $d \ge 4$) is always larger than the
1204: direct one, eq.\ \pref{eldir}, because $(m_e/m_w)^{d-2} \sim (6
1205: \times 10^{-6})^{d-2} \ll (\alpha_w /4 \pi) \sim 3\times 10^{-3}$.
1206: 
1207: \medskip
1208: \centerline{\epsfxsize=2.5in\epsfbox{electron.eps}} {\small Figure
1209: 4. Two-loop contribution to electron self-energy from W boson
1210: and graviton, which can dominate over the one-loop graph
1211: with gravitons alone, Fig.\ 2.}
1212: \medskip
1213: 
1214: The higher-loop result can also provide a larger estimate for
1215: $b_e$, to wit:
1216: %
1217: \eqa \label{elestb} b_e M_d^2 &\sim& \left( {\alpha_w \over 4 \pi}
1218: \right) \left( {\eps^2 \over (4 \pi)^{[d/2]} }
1219: \right) \left( {m_w \over M_d} \right)^{d-4} \qquad (d\ge 5)\\
1220: &\sim& \left\{ \matrix{ 4 \times 10^{-13} \, \left( {\eps \times
1221: 10^{3}} \right)^2 \, \left( { \hbox{TeV} / M_5} \right) & (d=5)
1222: \cr 9 \times 10^{-15} \, \left( {\eps \times 10^{3}} \right)^2 \,
1223: \left( { \hbox{TeV} / M_6} \right)^2 & (d=6) \cr} \right.  \, .\nn
1224: \eeqa
1225: %
1226: This dominates the direct one-loop result for all $d \ge 5$,
1227: because $(m_e/m_w)^{d-4} \sim (6 \times 10^{-6})^{d-4} \ll
1228: (\alpha_w /4 \pi) \sim 3\times 10^{-3}$.
1229: 
1230: When $d=4$ the direct one-loop result wins, for which the results
1231: of the previous section give
1232: %
1233: \eqa \label{elestbd4} b_e M_4^2 &=& - 12 \left( { \eps \over 8
1234: \pi } \right)^2 \qquad (d= 4) \\
1235: &=& -2 \times 10^{-8} \, \left( {\eps \times 10^3} \right)^2 .\nn\\
1236: \eeqa
1237: 
1238: 
1239: \subsection{Protons}
1240: %
1241: The direct calculation of section III applies directly for the
1242: proton only within the effective theory below a GeV, because it is
1243: only within this theory that the proton may be considered to be
1244: elementary. For higher scales the relevant degrees of freedom are
1245: quarks and gluons, for which we must estimate the size of
1246: Lorentz-violating effects.
1247: 
1248: The Lorentz-violating gluon and quark contributions may be estimated using our
1249: photon and electron results, eqs.~\pref{phest} and \pref{elest}, giving:
1250: %
1251: \eqa \label{glest}
1252: c_g^2 -1 &\sim& \left( {\alpha_s \over 4 \pi}
1253: \right) \left( {\eps \over (4 \pi)^{[d/2]}} \right) \; \left( {m_t
1254: \over M_d} \right)^{d-2} \nn\\
1255: c_q^2 -1 &\sim& \left( {\alpha_w \over 4 \pi}
1256: \right) \left( {\eps \over (4 \pi)^{[d/2]}} \right) \; \left(
1257: {m_w \over M_d} \right)^{d-2} \eeqa
1258: %
1259: where $\alpha_s$ is the QCD coupling, which we take to be $0.119$.
1260: From this we estimate the
1261: proton result to be
1262: %
1263: \eqa \label{prest}
1264: c_p^2 -1 &\sim& \hbox{Max}(c_g^2 -1,c_q^2 -1) \nn\\
1265: &\sim& \left( {\alpha_s \over 4 \pi}
1266: \right) \left( {\eps \over (4 \pi)^{[d/2]}} \right) \; \left( {m_t
1267: \over M_d} \right)^{d-2} \nn\\
1268: &\sim& \left\{ \matrix{ 2\times 10^{-9} \, \left( {\eps \times
1269: 10^{3}} \right) \, \left( { \hbox{TeV} / M_4} \right)^2 & (d=4)\cr
1270: 3 \times 10^{-10} \, \left( {\eps \times 10^{3}} \right) \, \left(
1271: { \hbox{TeV} / M_5} \right)^3 & (d=5) \cr
1272: 4 \times 10^{-12} \, \left( {\eps \times 10^{3}} \right) \, \left(
1273: { \hbox{TeV} / M_6} \right)^4 & (d=6) .\cr} \right. \nn \eeqa%
1274: 
1275: The loop contributions to $b_p$ can also dominate for large
1276: numbers of dimensions, but not for $d=4$. We have the higher-loop
1277: contribution
1278: %
1279: \eqa \label{prestb} b_p M_d^2 &\sim& \left( {\alpha_s \over 4 \pi}
1280: \right) \left( {\eps^2 \over (4 \pi)^{[d/2]} }
1281: \right) \; \left( {m_f \over M_d} \right)^{d-4} \qquad (d\ge 5)\\
1282: &\sim& \left\{ \matrix{ 1 \times 10^{-11} \, \left( {\eps \times
1283: 10^{3}} \right)^2 \, \left( { \hbox{TeV} / M_5} \right) & (d=5)
1284: \cr 1 \times 10^{-13} \, \left( {\eps \times 10^{3}} \right)^2 \,
1285: \left( { \hbox{TeV} / M_6} \right)^2 & (d=6) \cr} \right.  \, ,\nn
1286: \eeqa
1287: %
1288: when $m_f = m_t$.
1289: 
1290: By contrast, for $d=4$ it is the direct one-loop result which
1291: dominates. In this case we have a result for fermions which is
1292: largely insensitive to the fermion mass:
1293: %
1294: \eqa \label{prestbd4} b_f M_4^2 &=& - 12 \left( { \eps \over 8
1295: \pi } \right)^2 \qquad (d= 4) \\
1296: &=& -2 \times 10^{-8} \, \left( {\eps \times 10^3} \right)^2 \, .
1297: \nn \eeqa
1298: %
1299: For low-energy gravitons in the effective theory below the proton
1300: mass, the proton may be considered to be elementary and
1301: eq.~\pref{prestbd4} can be directly applied to the proton itself.
1302: 
1303: For higher-energy gravitons, eq.~\pref{prestbd4} would instead be
1304: applied to the light quarks, and the result for the proton would
1305: be obtained by taking the matrix element for the resulting
1306: effective quark operator, such as ${\cal O}_{\rm eff} = \eps^2 \,
1307: u^\mu u^\nu u^\alpha u^\beta \; \overline{q} \gamma_\mu
1308: \partial_\nu \partial_\alpha \partial_\beta \, q$, within the
1309: proton. On dimensional grounds this would produce the same result
1310: as eq.~\pref{prestbd4}, but potentially with a different numerical
1311: coefficient. For concreteness, when comparing with the observables
1312: we use eq.~\pref{prestbd4} including the numerical factor obtained
1313: for an elementary proton.
1314: 
1315: Notice that the above estimates suggest that so long as
1316: higher-loop contributions dominate, a hierarchy is to be expected:
1317: $| c_p^2 -1 | \gg | c_\gamma^2 - 1 | \sim | c_e^2 - 1|$. The same
1318: would {\it not} be expected to apply for $|b_p|, |b_\gamma|$ and
1319: $|b_e|$ when $d=4$, however.
1320: 
1321: \section{Experimental Constraints}
1322: %
1323: We now turn to the experimental constraints which can be imposed
1324: on the quantities $c^2 -1$ and $b$. Because we are interested in
1325: order-of-magnitude bounds, we derive constraints on each of these
1326: quantities as if they had arisen in the absence of the other. That
1327: is, we consider bounds on $c^2 -1$ while taking $b =0$ and vice
1328: versa. This is justified unless there is an unnatural cancellation
1329: between the contributions of these quantities to physical
1330: observables.
1331: 
1332: \subsection{Existing Bounds on $c_f^2 - c_\gamma^2$}
1333: %
1334: Good bounds exist on a difference between the propagation speeds
1335: of fermions and photons. Among those which do not depend on the
1336: sign of $c_f -c_\gamma$ are \cite{Glashow}:
1337: %
1338: \eqa |c_f - c_\gamma | &<& 6 \times 10^{-22} \qquad \hbox{Atomic
1339: spectroscopy} \nn\\
1340: |c' - c|_{\mu e} &<& 4 \times 10^{-21} \qquad \mu \to e\gamma
1341: \nn\\
1342: |c_{\sss KL} - c_{\sss KS}| &<& 3 \times 10^{-21} \qquad K-\overline{K}
1343: \quad\hbox{oscillations} . \eeqa
1344: %
1345: In the second bound $c'$ denotes a particular combination of the
1346: Lorentz-violating couplings in flavor space, whose details are
1347: not important in what follows.
1348: 
1349: Although these bounds apply to both signs of $c^2 -c_\gamma^2$,
1350: they are also subject to specific assumptions which need not apply
1351: to our calculation. For instance the $\mu\to e\gamma$ bound
1352: assumes the existence of Lorentz-violating terms which also
1353: violate lepton number, while the bounds involving neutral kaons
1354: require strangeness violation in addition to Lorentz-violation.
1355: Since the loops which would produce both type of flavor-symmetry
1356: violations from Lorentz-violation in the graviton sector are
1357: further suppressed by masses and mixing angles, we do not consider
1358: these bounds further.
1359: 
1360: The bounds from atomic spectroscopy are usually derived under the
1361: assumption that all matter particles share the same maximum
1362: propagation speed which differs from that of the photon -- {\it
1363: i.e.,} $c_f \ne c_\gamma$ is independent of $f$. Nevertheless
1364: these bounds are likely also to apply if $c_f$ differs for
1365: electrons and nucleons. Although we use the bound of
1366: ref.~\cite{Glashow} in what follows, we regard a more careful
1367: analysis of constraints which are implied by these experiments to
1368: be worth pursuing.
1369: 
1370: If we apply these bounds directly to protons, we find $|c_p^2 -
1371: c_\gamma^2| < 6 \times 10^{-22}$. Using our previous estimates for
1372: the proton, eq.~\pref{prest}, and our expectation (see above) that
1373: $c_p^2$ differs from unity by more than does $c_\gamma^2$, we find
1374: comparison with the bound implies,
1375: %
1376: \eq \eps < \left( {4 \pi \over \alpha_s} \right)
1377: \; (4 \pi)^{[d/2]} \, \left( {M_d \over m_t} \right)^{d-2}
1378: 10^{-21}. \eeq
1379: 
1380: This is stronger than the direct bound $\eps
1381: <10^{-6}$\footnote{Using the weaker bound $\eps < 10^{-3}$ makes
1382: the limits for $M_d$ larger by a factor of $10^{3/(d-2)}$.} only
1383: if
1384: %
1385: \eq {M_d \over \hbox{TeV}} < 0.2 \; \left[ {10^{13} \over (4
1386: \pi)^{[d/2]}} \right]^{1/(d-2)} \sim \left\{ \matrix{3\times 10^4
1387: & (d=4) \cr 700 & (d=5) \cr 46 & (d=6) } \right. .\eeq
1388: %
1389: The bound so obtained is not useful for $d=4$. With extra
1390: dimensional gravitons the bound becomes useful only when $M_d$ is
1391: very small. It is only of borderline interest for the ADD
1392: scenario, for which $d=6$ but $M_6 > 50$ TeV is required from
1393: stellar-cooling bounds \cite{ADDBounds}. (We do not quote here the
1394: stronger limits coming from the non-observance of gamma-ray decay
1395: products \cite{ADDGBounds}, or which apply for supersymmetric
1396: models \cite{LED}, because these may be evaded depending on
1397: model-dependent details.)
1398: 
1399: The bound {\it is} competitive, however, for the warped 5D model
1400: described below eq.\ (\ref{bfeqs}), where $M_5$ is effectively the
1401: TeV scale.
1402: 
1403: \subsection{\v Cerenkov Bounds on $b_f$ and $c_f^2 -1$}
1404: 
1405: We next consider constraints from high-energy cosmic rays,
1406: typically protons or photons.  Their observation precludes the existence of
1407: processes which would too-efficiently deplete the energy of these
1408: particles. Since the particles involved are at higher energies, these
1409: processes are more sensitive to the higher-momentum coefficients,
1410: $b_f, b_\gamma$, than are low-energy laboratory limits.
1411: 
1412: Lorentz-violation can introduce dangerous processes by allowing decays
1413: in vacuum of the form $p \to p\gamma$ or $\gamma \to e^+ e^-$ or
1414: $\gamma \to p \overline{p}$. Such decays are precluded by energy
1415: and momentum conservation in Lorentz-invariant systems, but {\it
1416: are} allowed given Lorentz-violating dispersion relations. For
1417: instance, the process $p \to p \gamma$ becomes allowed if the
1418: dispersion relation raises the energy of the proton at a given
1419: momentum more than it does for the photon. Similarly, the process
1420: $\gamma \to e^+ e^-$ can occur if the photon energy is raised more
1421: than the electron's at a given momentum.
1422: 
1423: As applied to changes in the maximum propagation speed, the
1424: resulting bounds are therefore one-sided, in the sense that they
1425: are only relevant for one sign of $c_p - c_\gamma$ or $c_\gamma -
1426: c_e$. For protons the strong one-sided bound obtained from $p \to
1427: p\gamma$ requires $c_p > c_\gamma$. Similarly, for $\gamma \to e^+
1428: e^-$ or $\gamma \to p \overline{p}$, $c_e < c_\gamma$ and $c_p <
1429: c_\gamma$ are required to use the one-sided bound.
1430: 
1431: Keeping in mind that our estimates of the previous sections imply
1432: $|\delta c_p^2| \gg |\delta c_\gamma^2 | \sim | \delta c_e^2 |$, we
1433: find the limits obtained in this way for electrons and protons are
1434: %
1435: \eqa -2 \times 10^{-8} < c_p^2 - c_\gamma^2 \sim c_p^2 -1 &<& 2 \times 10^{-23} \nn\\
1436: c_e^2 - c_\gamma^2 &>& -6 \times 10^{-15}. \eeqa
1437: %
1438: In order of magnitude, these bounds are obtained by demanding that
1439: $|c_f^2 -c_\gamma^2 |< m_f^2/E^2$ where $m_f$ is the relevant
1440: fermion mass, $E$ is the energy of the observed cosmic ray, and
1441: the bound only applies to the appropriate sign of $c_f^2 -
1442: c_\gamma^2$. The bound from $\gamma \to f \overline{f}$ is weaker
1443: than that from $p \to p \gamma$ because the most energetic cosmic
1444: ray proton has $E \sim 10^{8}$ TeV, while the most energetic
1445: observed gamma ray has $E \sim 50$ TeV.
1446: 
1447: A similar bound also applies to the parameter $b_f$, which
1448: controls the dispersive part of the dispersion relation, eq.\
1449: \pref{disprel}. For $p \to p\gamma$ the bound then applies only if
1450: $b_p > b_\gamma$ and for $\gamma \to f \overline{f}$ it requires
1451: $b_f < b_\gamma$. When the bound applies, it is of order $|b_f -
1452: b_\gamma | < m_f^2/E^4$.
1453: 
1454: We find in this way the constraints
1455: %
1456: \eqa -(5 \times 10^{8} \; \hbox{GeV})^{-2} < &b_p - b_\gamma& < (3
1457: \times 10^{22} \;
1458: \hbox{GeV})^{-2} \nn\\
1459: &b_e - b_\gamma & > - (1 \times 10^{12} \; \hbox{GeV})^{-2} .
1460: \eeqa
1461: 
1462: \subsubsection{$b_p > b_\gamma$}
1463: Consider first the case $b_p > b_\gamma$, for which the best bound
1464: applies. For $d>4$ we compare this to the order-of-magnitude of
1465: the higher-loop results, with the resulting expectation $|b_p| \gg
1466: |b_\gamma|$. Using eq.~\pref{prestb} for $b_p$, as discussed in
1467: the previous section, and assuming the sign of the result is the
1468: one relevant for the bound's applicability, we obtain the limit
1469: %
1470: \eq \eps < \left[ {(4 \pi)^{[(d+2)/2]}\over{\alpha_s}} \, \left( {M_d \over \ m_t}
1471: \right)^{(d-4)} \right]^{1/2} \; \left( { M_d \over 3 \times
1472: 10^{22} \; \hbox{GeV} } \right). \eeq
1473: 
1474: This is an improvement over $\eps < 10^{-6}$ if
1475: %
1476: \eqa {M_d \over \hbox{TeV}} &<& \left[ { 3 \times 10^{13} \,  \ (
1477: {0.175})^{(d-4)/2} \sqrt{{\alpha_s}\over (4 \pi)^{[d/2]+1}}}
1478: \right]^{2/(d-2)} \nn\\
1479: &<& \left\{ \matrix{ 2 \times 10^{7}&(d=5) \cr 9 \times
1480: 10^4&(d=6)\cr 6 \times 10^4 & (d=7) \cr 800 &(d=8)\cr 300
1481: &(d=9)\cr 70&(d=10) \, .} \right. \eeqa
1482: %
1483: (An improvement on the bound $\eps < 10^{-3}$ is obtained for
1484: $M_d$ which can be a factor $10^{6/(d-2)}$ larger.) We see the
1485: bound can be competitive for all $d$, provided $M_d$ is in the
1486: lower end of its allowed range. It is particularly strong for the
1487: ADD ($d=6$) and warped RS ($d=5$) cases, for which $M_d$ is in the
1488: TeV range.
1489: 
1490: To apply the bound to $d=4$ we instead use the one-loop result,
1491: for which the numerical factors and sign are known from our
1492: previous calculation (provided that the high-energy quark
1493: contribution does not dominate that for the low-energy proton, as
1494: discussed above). Although $b_p$ which is obtained is negative,
1495: the bound nonetheless applies because for $d=4$ we know that
1496: $b_\gamma$ is also negative, with $b_f > b_\gamma$ ({\it c.f.}
1497: eqs.~\pref{phestbd4} and \pref{prestbd4}).%
1498: \footnote{
1499: In fact, at the energies under consideration the graviton should resolve
1500: the proton as a number of partons, each carrying a small momentum
1501: fraction $x \ll 1$.  Since the importance of the
1502: dispersive term grows with momentum as
1503: $p^2$, there will be a suppression $\sim x_{\rm av}^2$.  Hence,
1504: compared to the photon, the effective $b$ for the proton may be close to
1505: zero.}
1506: We find
1507: %
1508: \eq \label{d4bound} b_p - b_\gamma = 10^{-8} \, {(\eps \times
1509: 10^3)^2 \over M_4^2} < (3 \times 10^{22} \; \hbox{GeV})^{-2} \eeq
1510: %
1511: which, using $M_4 = 2 \times 10^{18}$ GeV, implies $\eps < 7
1512: \times 10^{-4}$ --- a result which is competitive with the
1513: terrestrial bound, $\eps < 10^{-3}$, when $\eps > 0$, but is not
1514: as good as the bound $\eps \lsim 10^{-6}$ obtained from the
1515: conservation of angular momentum of the sun (see the footnote on
1516: page 5).
1517: 
1518: \subsubsection{$b_p < b_\gamma$}
1519: If $b_p < b_\gamma$, then the best bound comes from the process
1520: $\gamma \to p\overline{p}$, which does not give a bound even as
1521: good as $\eps < 10^{-3}$ for any $d$ given the constraint $M_d >
1522: 1$ TeV.
1523: 
1524: \subsection{$b_e < b_\gamma$}
1525: 
1526: The bound when $b_e < b_\gamma$ is not strong enough to be
1527: interesting for $d=4$, so we consider only higher dimensions.
1528: Since the order-of-magnitude result for the electron has $|b_e|
1529: \sim |b_\gamma|$ we have
1530: %
1531: \eq b_\gamma \sim {\eps^2  \over {(4 \pi)^{[d/2]} \, M_d^2}} \;
1532: \left( {m_t \over M_d} \right)^{d-4}\; \,\left({\alpha \over
1533: 4\pi}\right),\eeq
1534: %
1535: and so, if $b_\gamma - b_e > 0$ the bound becomes:
1536: %
1537: \eq \eps < \left[ {(4 \pi)^{[d/2]+1}\over {\alpha}} \left( {M_d
1538: \over \ m_t} \right)^{(d-4)} \right]^{1/2} \; \left( { M_d \over
1539: 10^{12} \; \hbox{GeV} } \right). \eeq
1540: %
1541: For $M_d > 1$ TeV this is never better than $\eps < 10^{-6}$, but
1542: it represents progress relative to $\eps < 10^{-3}$ if
1543: %
1544: \eqa {M_d \over \hbox{TeV}} &<& \left[ { \, 10^{6} \ (
1545: {0.175})^{(d-4)/2} \sqrt{\alpha \over (4 \pi)^{[d/2]+1}}} \,
1546: \right]^{2/(d-2)} \\
1547: &=&  \left\{\matrix{ 90 & (d=5)\cr 10 &(d=6)} \right. \, ,\eeqa
1548: %
1549: which is only of interest for 5D warped scenarios.
1550: 
1551: \subsection{Direct Bounds on $b_\gamma$ }
1552: The observation of cosmologically distant gamma rays from
1553: gamma-ray bursters can provide a bound on $|b_\gamma|$ which is
1554: similar in strength to the one just obtained, but applies equally
1555: well to both signs of the result. Ref.~\cite{Ellis} argues that
1556: the absence of dispersion in these signals provides a sensitivity
1557: to changes in the photon dispersion which can be as small as
1558: %
1559: \eq |b_\gamma| < {( 9\times 10^{11} \; \hbox{GeV})^{-2}},\eeq
1560: %
1561: which would correspond to
1562: %
1563: \eq \eps < \left[ {(4 \pi)^{[d/2]+1}\over{\alpha}} \, \left( {M_d
1564: \over \ m_t} \right)^{(d-4)} \right]^{1/2} \; \left( { M_d \over
1565: 10^{11} \; \hbox{GeV} } \right). \eeq
1566: 
1567: This is better than $\eps < 10^{-3}$ provided
1568: %
1569:  \eqa {M_d \over \hbox{TeV}} &<& \left[ { 10^{5} \,  \ (
1570: {0.175})^{(d-4)/2} \sqrt{{\alpha}\over (4 \pi)^{[d/2]+1}}}
1571: \right]^{2/(d-2)} \nn\\
1572: &=& \left\{ \matrix{ 20 &(d=5) \cr 3 &(d=6)} \right. \eeqa
1573: %
1574: which is comparable in strength to what was found above.
1575: 
1576: \section{Conclusions}
1577: %
1578: Brane-world scenarios allow the possibility of potentially strong
1579: preferred-frame effects arising from extra-dimensional bulk
1580: physics, and suggest these effects may be limited to the graviton
1581: sector. Depending on their sign, the magnitude of such changes to
1582: graviton dispersion can be comparatively large because of the
1583: graviton's extremely weak interactions. Motivated by this
1584: observation, we have explored how gravitational Lorentz-violation
1585: in a brane-world picture influences the properties of observable
1586: particles. We obtain the following results:
1587: 
1588: \bigskip
1589: 
1590: {\it 1.} We analyze how Lorentz violation in extra dimensions
1591: arises within the low-energy four-dimensional field theory which
1592: is obtained once the extra dimensions are integrated out. We find
1593: the leading contributions to be the appearance of potentially
1594: different metrics in the kinetic terms which govern the
1595: propagation of the various low-energy particles. If
1596: preferred-frame effects dominate violations of translation or
1597: rotation invariance (in the preferred frame) then in flat space
1598: the metric seen by particle type `$k$' may be written
1599: %
1600: \eq G_{\mu\nu} = \eta_{\mu\nu} + (1 - c_k^2) \, u_\mu u_\nu \eeq
1601: %
1602: where $u^\mu$ is the (approximately constant) 4-velocity which
1603: defines the preferred frame. $c_k$ may be interpreted as the
1604: maximum propagation speed of this particle type.
1605: 
1606: {\it 2.} If subleading contributions at low energies are also
1607: included, then more complicated changes to the dispersion relation
1608: become possible. For low momenta these dispersion relations become
1609: $E_k^2 = p^2 c_k^2 + b_k p^4 + \dots$ in the preferred frame. The
1610: next-to-leading coefficient $b_k$ causes dispersive propagation if
1611: it is nonzero. (Odd powers of $p$, like $p^3$, typically do not
1612: arise in the dispersion relation since they are usually forbidden
1613: by selection rules like rotation invariance in the preferred
1614: frame.)
1615: 
1616: {\it 3.} Strong observational constraints exist which preclude
1617: large differences between the values $c_k$ and $b_k$ for photons
1618: and other particles. The strongest of these come from the absence
1619: of a dependence on the Earth's velocity through space in atomic
1620: spectra, and from the absence of \v Cerenkov-like decays of
1621: very-high-energy cosmic rays.
1622: 
1623: {\it 4.} We compute how graviton loops can bring the news of
1624: Lorentz violation in the graviton sector to other particles for
1625: which stronger constraints exist. We do so quite generally at low
1626: energies for the purely massless 4D graviton (the lowest KK mode).
1627: We also compute these loops for the entire KK tower of gravitons,
1628: in the approximation that the dominant contribution comes from
1629: gravitons whose wavelengths are much shorter than are the typical
1630: curvature scales of the extra-dimensional metric.
1631: 
1632: {\it 5.} We find that one-loop contributions for photons from
1633: graviton-induced Lorentz violation are small. Keeping the finite
1634: and log-divergent parts, we find gravitons do not induce any
1635: change at all in the photon maximum propagation speed, $c_\gamma$.
1636: The purely 4D graviton induces a dispersive term, $b_\gamma \sim
1637: (c_g - 1)^2 /M_p^2$. The contribution of the rest of the KK
1638: graviton tower to the photon energy tends to be suppressed by
1639: further powers of the photon momenta, and so is not important at
1640: low energies. Unfortunately this makes the strong constraints on
1641: photon properties based on gamma-ray bursts and dispersion in
1642: quasar signals \cite{Ellis} largely irrelevant to this kind of
1643: Lorentz violation.
1644: 
1645: {\it 6.} Using the same approximations, we find fermions acquire
1646: changes to their low-energy dispersion relations with an amount
1647: which varies strongly with the fermion's mass. Among the three
1648: most abundant particles in everyday life -- electrons, protons and
1649: photons -- this predicts by far the biggest effects for protons.
1650: Contributions to $c_k -1 $ are linear in $c_g -1 $ (if $c_g$ is
1651: the graviton maximum speed) while those to $b_k$ are quadratic in
1652: $c_g - 1$.
1653: 
1654: {\it 7.} The strong mass dependence makes the results very
1655: sensitive to the high-energy spectrum of the theory, since heavy
1656: particles embedded in higher loops can produce larger
1657: contributions to low-energy Lorentz-violating effects than do the
1658: direct graviton loops. An estimate of this effect using the top
1659: quark or $W$-boson as the heavy particle suggests that protons
1660: receive larger contributions than do photons or electrons.
1661: 
1662: {\it 8.} We find that current atomic constraints on $c_k - 1$ for
1663: observable particles can provide limits on $c_g - 1$ which are
1664: competitive with the direct bound $c_g - 1 \lsim 10^{-6}$ arising
1665: from post-Newtonian corrections in the solar system, but only if
1666: $d=5$ or 6 and if the higher-dimensional Planck mass, $M_d$, is as
1667: close as possible to the TeV scale. For instance, for a warped
1668: $d=5$ model with $M_5 \sim 10$ TeV, we find $|c_g - 1| < 3 \times
1669: 10^{-15}\, (M_5/\hbox{TeV})^3$. For $c_g > 1$ this is an
1670: improvement over the direct bound $c_g - 1 < 10^{-6}$ provided
1671: $M_5 < 700$ TeV.
1672: 
1673: {\it 8.} We find that stronger constraints on $c_g - 1$ arise from
1674: limits on $b_k - b_\gamma$, depending on the dimension of the
1675: extra-dimensional spacetime, and on how low the $d$-dimensional
1676: Planck scale, $M_d$, is. Lower values of $M_d$ lead to better
1677: bounds, which can go up to values of order $M_d \sim 10^7$ TeV
1678: (for $d=5$), provided that $b_p > b_\gamma$. The sign of this
1679: quantity is important, because the bounds which are most
1680: constraining are those which are based on the absence of
1681: too-efficient energy-loss mechanisms for the highest energy cosmic
1682: rays (which we take to be protons), and the decay channel $p \to
1683: p\gamma$ is only open if $b_p > b_\gamma$. Strikingly, purely
1684: 4-dimensional graviton loops can give contributions with the right
1685: sign, and which are large enough to produce bounds which are of
1686: order $\eps < 7 \times 10^{-4}$. This makes them comparable with
1687: those from obtained from terrestrial experiments and the binary
1688: pulsar.
1689: 
1690: \bigskip
1691: 
1692: Perhaps our most surprising result is that, in some regimes,
1693: graviton loops are already being constrained by observational
1694: data. This is yet another striking way in which the brane-world
1695: picture can run against pre-brane-world intuition.
1696: 
1697: \vspace{1 cm}
1698: \begin{center}
1699: {\bf Acknowledgements}
1700: \end{center}
1701: 
1702: We would like to acknowledge fruitful conversations with Debajyoti Chauduri, 
1703: Maxim Pospelov and Xerxes Tata. C.B. wishes to acknowledge
1704: discussions with Detlef Nolte and Witold Skiba at an early point
1705: in this research. J.M. acknowledges discussions with A.Pomarol.
1706: The research of C.B., J.C. and E.F. has been
1707: supported in part by N.S.E.R.C. of Canada and F.C.A.R. of
1708: Qu\'ebec. J.M acknowledges  support by CICYT Research Project
1709: AEN99-0766, MCyT  and the Theory Group of McGill
1710: University.
1711: 
1712: \section{Appendix A}
1713: In this appendix we record the expressions for the Feynman rule
1714: for the photon-graviton vertex, as obtained from the
1715: Einstein-Maxwell action, eq.\ \pref{DEHAction}. One finds:
1716: %
1717: \eqa \label{vertices} V^{\alpha\beta:\mu\nu}(P,Q) &=& (P^\mu Q^\nu
1718: + P^\nu Q^\mu) \, \eta^{\alpha\beta} \nn\\
1719: && \quad + (P^\beta Q^\alpha - P\cdot Q \, \eta^{\alpha \beta}) \,
1720: \eta^{\mu\nu} \nn\\
1721: && \quad - (P^\mu Q^\alpha - P\cdot Q \, \eta^{\mu\alpha} ) \,
1722: \eta^{\nu \beta} \nn\\
1723: && \quad - ( P^\beta Q^\mu - P\cdot Q \, \eta^{\mu\beta} ) \,
1724: \eta^{\nu\alpha} \nn\\
1725: && \quad - P^\nu Q^\alpha \, \eta^{\mu\beta} - P^\beta Q^\nu \,
1726: \eta^{\mu\alpha} , \eeqa
1727: %
1728: which has the required symmetry properties:
1729: %
1730: \eq V^{\alpha\beta:\mu\nu}(P,Q) = V^{\alpha\beta:\nu\mu}(P,Q) =
1731: V^{\beta\alpha:\mu\nu}(Q,P). \eeq
1732: 
1733: \section{Appendix B}
1734: We here record in more detail the results of the fermion
1735: self-energy calculation, using a graviton loop in $d$ dimensions.
1736: 
1737: \subsection{d=5}
1738: %
1739: For five spacetime dimensions evaluation of the graviton loop
1740: gives
1741: %
1742: \eqa \scA_5 &=& - m^4 \lambda_5 \; \left(\frac{25}{9} + \frac{74}{9}
1743: \, \eps + {85\over 9} \, \eps^2 + ...\right) ,\nn\\
1744:  \scB_5 &=&  - m^3 \lambda_5 \; \left(\frac{35}{18} \, \eps + \frac{839}{180} \, \eps^2 + ...\right) ,\nn\\
1745:  \scC_5 &=& - m^2 \lambda_5  \; \left(\frac{25}{6}\, \eps + \frac{517}{60}\, \eps^2 + ...\right) ,\nn\\
1746:  \scD_5 &=& - m \lambda_5 \; \left(\frac{101}{30} \, \eps^2 + ...\right) ,\nn\\
1747:  \scE_5 &=& -\lambda_5 \; \left(\frac{29}{30} \, \eps^2 + ... \right)\; . \eeqa
1748: %
1749: where $\lambda_5 = (\kappa_5/8\pi)^2$. The result is finite in
1750: dimensional regularization because one-loop results in odd
1751: dimensions always are with this regularization scheme. This result
1752: agrees with the finite part as computed by directly cutting off
1753: the momentum integrals and ignoring the divergent terms (none of
1754: which are logarithmically divergent).
1755: 
1756: These lead to the dispersion relation of eq.\ \pref{disprel}:
1757: %
1758: \eqa c_f^2 -1 &=&  m_f^3 \lambda_5  \; \left[ \frac{110}{9} \,
1759: \eps + \frac{239}{9} \, \eps^2 + ... \right], \nn\\
1760: b_f &=&   m_f \, \lambda_5 \left[ \frac{26}{3} \, \eps^2 + ...
1761: \right] \eeqa
1762: %
1763: Notice that both of these results are positive (provided $\eps >
1764: 0$).
1765: 
1766: \subsection{d=6}
1767: %
1768: For six spacetime dimensions the graviton loop gives
1769: %
1770: \eqa \scA_6 &=& - m^5 \lambda_6 \; \left(\frac{9}{4} + \frac{25}{4}
1771: \, \eps + {53\over 8} \, \eps^2 + ...\right) ,\nn\\
1772:  \scB_6 &=& - m^4 \lambda_6 \; \left(\frac{6}{5} \, \eps + \frac{83}{30} \, \eps^2 + ...\right) ,\nn\\
1773:  \scC_6 &=& - m^3 \lambda_6  \; \left(\frac{18}{5}\, \eps + \frac{107}{15}\, \eps^2 + ...\right) ,\nn\\
1774:  \scD_6 &=& - m^2 \lambda_6 \; \left(\frac{34}{15} \, \eps^2 + ...\right) ,\nn\\
1775:  \scE_6 &=& - m \lambda_6 \; \left(\frac{4}{3} \, \eps^2 + ... \right)\; . \eeqa
1776: %
1777: where $\lambda_6 = 2\kappa_6^2/(8\pi)^3 \, {\cal L}$. As before
1778: ${\cal L} = \log(\Lambda^2/\mu^2) = 2/(6-n)$, when evaluated with
1779: an ultraviolet cutoff, $\Lambda$, and in dimensional
1780: regularization.
1781: 
1782: These lead to the $d=6$ results:
1783: %
1784: \eqa c_f^2 -1 &=&  m_f^4 \, \lambda_6  \; \left[ \frac{48}{5}
1785: \, \eps + \frac{99}{5} \, \eps^2 + ... \right], \nn\\
1786: b_f &=&  m_f^2 \, \lambda_6 \left[ \frac{36}{5} \, \eps^2 + ...
1787: \right] \eeqa
1788: %
1789: Again both results are positive for $\eps > 0$.
1790: 
1791: \subsection{d=7}
1792: %
1793: Next, seven spacetime dimensions:
1794: %
1795: \eqa \scA_7 &=&  m^6 \lambda_7 \; \left(\frac{49}{25} +
1796: \frac{26}{5} \, \eps + {129\over 25} \, \eps^2 + ...\right) ,\nn\\
1797:  \scB_7 &=&  m^5 \lambda_7 \; \left(\frac{21}{25} \, \eps + \frac{131}{70} \, \eps^2 + ...\right) ,\nn\\
1798:  \scC_7 &=&  m^4 \lambda_7  \; \left(\frac{49}{15}\, \eps + \frac{263}{42}\, \eps^2 + ...\right) ,\nn\\
1799:  \scD_7 &=&  m^3 \lambda_7 \; \left(\frac{59}{35} \, \eps^2 + ...\right) ,\nn\\
1800:  \scE_7 &=&  m^2 \lambda_7 \; \left(\frac{53}{35} \, \eps^2 + ... \right)\; . \eeqa
1801: %
1802: where $\lambda_7 = \kappa_7^2/[6(4\pi)^3]$, and is finite when
1803: evaluated in dimensional regularization.
1804: 
1805: These lead to the $d=7$ results:
1806: %
1807: \eqa c_f^2 -1 &=& - m_f^5 \, \lambda_7  \; \left[ \frac{616}{75}
1808: \, \eps + \frac{244}{15} \, \eps^2 + ... \right], \nn\\
1809: b_f &=& - m_f^3 \, \lambda_7 \left[ \frac{32}{5} \, \eps^2 + ...
1810: \right] \eeqa
1811: %
1812: Here both results are negative for $\eps > 0$.
1813: 
1814: \subsection{d=8}
1815: %
1816: For eight spacetime dimensions we have
1817: %
1818: \eqa \scA_8 &=&  m^7 \lambda_8 \; \left(\frac{16}{9} +
1819: \frac{41}{9}
1820: \, \eps + {77\over 18} \, \eps^2 + ...\right) ,\nn\\
1821:  \scB_8 &=&  m^6 \lambda_8 \; \left(\frac{40}{63} \, \eps + \frac{173}{126} \, \eps^2 + ...\right) ,\nn\\
1822:  \scC_8 &=&  m^5 \lambda_8  \; \left(\frac{64}{21}\, \eps + \frac{239}{42}\, \eps^2 + ...\right) ,\nn\\
1823:  \scD_8 &=&  m^4 \lambda_8 \; \left(\frac{4}{3} \, \eps^2 + ...\right) ,\nn\\
1824:  \scE_8 &=&   m^3 \lambda_8 \; \left(\frac{34}{21} \, \eps^2 + ... \right)\; . \eeqa
1825: %
1826: where $\lambda_8 = 2\kappa_8^2/(8\pi)^4 \, {\cal L}$. As before
1827: ${\cal L} = \log(\Lambda^2/\mu^2) = 2/(8-n)$, when evaluated with
1828: an ultraviolet cutoff, $\Lambda$, and in dimensional
1829: regularization.
1830: 
1831: These lead to the $d=8$ results:
1832: %
1833: \eqa c_f^2 -1 &=& - m_f^6 \, \lambda_8  \; \left[ \frac{464}{63}
1834: \, \eps + \frac{890}{63} \, \eps^2 + ... \right], \nn\\
1835: b_f &=& -  m_f^4 \, \lambda_8 \left[ \frac{124}{21} \, \eps^2 + ...
1836: \right] \eeqa
1837: %
1838: Here both results are negative for $\eps > 0$.
1839: 
1840: \subsection{d=9}
1841: %
1842: For $d=9$ we have
1843: %
1844: \eqa \scA_9 &=&  -\,m^8 \lambda_9 \; \left(\frac{81}{49} +
1845: \frac{202}{49}
1846: \, \eps + {181\over 49} \, \eps^2 + ...\right) ,\nn\\
1847:  \scB_9 &=& -\, m^7 \lambda_9 \; \left(\frac{99}{196} \, \eps +
1848:  \frac{3755}{3528} \, \eps^2 + ...\right) ,\nn\\
1849:  \scC_9 &=& -\, m^6 \lambda_9  \; \left(\frac{81}{28}\, \eps +
1850:  \frac{2665}{504}\, \eps^2 + ...\right) ,\nn\\
1851:  \scD_9 &=& -\, m^5 \lambda_9 \; \left(\frac{277}{252} \, \eps^2 + ...\right) ,\nn\\
1852:  \scE_9 &=& -\, m^4 \lambda_9 \; \left(\frac{425}{252} \, \eps^2 + ... \right)\; . \eeqa
1853: %
1854: where $\lambda_9 = \kappa_9^2/[15 (4\pi)^4]$.
1855: 
1856: These lead to the results:
1857: %
1858: \eqa c_f^2 -1 &=&  m_f^7 \, \lambda_9  \; \left[ \frac{333}{49}
1859: \, \eps + \frac{1245}{98} \, \eps^2 + ... \right], \nn\\
1860: b_f &=&  m_f^5 \, \lambda_9 \left[ \frac{39}{7} \, \eps^2 + ...
1861: \right] \eeqa
1862: %
1863: Here both results are positive for $\eps > 0$.
1864: 
1865: \subsection{d=10}
1866: %
1867: Finally, for ten spacetime dimensions:
1868: %
1869: \eqa \scA_{10} &=& -\,m^9 \lambda_{10} \; \left(\frac{25}{16} +
1870: \frac{61}{16}
1871: \, \eps + {105\over 32} \, \eps^2 + ...\right) ,\nn\\
1872:  \scB_{10} &=& -\,m^8 \lambda_{10} \; \left(\frac{5}{12} \, \eps +
1873:  \frac{103}{120} \, \eps^2 + ...\right) ,\nn\\
1874:  \scC_{10} &=& -\, m^7 \lambda_{10}  \; \left(\frac{25}{9}\, \eps +
1875:  \frac{449}{90}\, \eps^2 + ...\right) ,\nn\\
1876:  \scD_{10} &=& -\, m^6 \lambda_{10} \; \left(\frac{14}{15} \, \eps^2
1877:  + ...\right) ,\nn\\
1878:  \scE_{10} &=& -\, m^5 \lambda_{10} \; \left(\frac{26}{15} \, \eps^2
1879:  + ... \right)\; . \eeqa
1880: %
1881: where $\lambda_{10} = 4\kappa_{10}^2/[3(8\pi)^5] \, {\cal L}$, and
1882: ${\cal L} = \log(\Lambda^2/\mu^2) = 2/(10-n)$, when
1883: evaluated with an ultraviolet cutoff, $\Lambda$, and in
1884: dimensional regularization.
1885: 
1886: These lead to the $d=10$ results:
1887: %
1888: \eqa c_f^2 -1 &=&  m_f^8 \, \lambda_{10} \; \left[
1889: \frac{115}{18} \, \eps + \frac{421}{36} \, \eps^2 + ... \right], \nn\\
1890: b_f &=&  m_f^6 \, \lambda_{10} \left[ \frac{16}{3} \, \eps^2 +
1891: ... \right] \eeqa
1892: %
1893: $\eps > 0$ ensures that both results in this case are positive.
1894: 
1895: 
1896: 
1897: \begin{thebibliography}{99}
1898: 
1899: \bibitem{HW}
1900: P.~Horava and E.~Witten,
1901: %``Eleven-Dimensional Supergravity on a Manifold with Boundary,''
1902: Nucl.\ Phys.\ B {\bf 475}, 94 (1996)
1903: [hep-th/9603142];
1904: %%CITATION = HEP-TH 9603142;%%
1905: P.~Horava and E.~Witten,
1906: %``Heterotic and type I string dynamics from eleven dimensions,''
1907: Nucl.\ Phys.\ B {\bf 460}, 506 (1996)
1908: [hep-th/9510209].
1909: %%CITATION = HEP-TH 9510209;%%
1910: 
1911: \bibitem{ADD}
1912: I.~Antoniadis,
1913: %``A Possible New Dimension At A Few Tev,''
1914: Phys.\ Lett.\ B {\bf 246}, 377 (1990);
1915: %%CITATION = PHLTA,B246,377;%%
1916: J.~D.~Lykken,
1917: %``Weak Scale Superstrings,''
1918: Phys.\ Rev.\ D {\bf 54}, 3693 (1996)
1919: [hep-th/9603133];
1920: %%CITATION = HEP-TH 9603133;%%
1921: I.~Antoniadis, N.~Arkani-Hamed, S.~Dimopoulos and G.~R.~Dvali,
1922: %``New dimensions at a millimeter to a Fermi and superstrings at a TeV,''
1923: Phys.\ Lett.\ B {\bf 436}, 257 (1998)
1924: [hep-ph/9804398];
1925: %%CITATION = HEP-PH 9804398;%%
1926: N.~Arkani-Hamed, S.~Dimopoulos and G.~R.~Dvali,
1927: %``The hierarchy problem and new dimensions at a millimeter,''
1928: Phys.\ Lett.\ B {\bf 429}, 263 (1998)
1929: [hep-ph/9803315];
1930: %%CITATION = HEP-PH 9803315;%%
1931: N.~Arkani-Hamed, S.~Dimopoulos and G.~R.~Dvali,
1932: %``Phenomenology, astrophysics and cosmology of theories with  sub-millimeter dimensions and TeV scale quantum gravity,''
1933: Phys.\ Rev.\ D {\bf 59}, 086004 (1999)
1934: [hep-ph/9807344].
1935: %%CITATION = HEP-PH 9807344;%%
1936: 
1937: \bibitem{OtherScales}
1938: E.~Witten,
1939: %``Strong Coupling Expansion Of Calabi-Yau Compactification,''
1940: Nucl.\ Phys.\ B {\bf 471}, 135 (1996)
1941: [hep-th/9602070];
1942: %%CITATION = HEP-TH 9602070;%%
1943: K.~Benakli,
1944: %``Phenomenology of low quantum gravity scale models,''
1945: Phys.\ Rev.\ D {\bf 60}, 104002 (1999)
1946: [hep-ph/9809582];
1947: %%CITATION = HEP-PH 9809582;%%
1948: C.~P.~Burgess, L.~E.~Iba\~nez and F.~Quevedo,
1949: %``Strings at the intermediate scale or is the Fermi scale dual to the  Planck scale?,''
1950: Phys.\ Lett.\ B {\bf 447}, 257 (1999)
1951: [hep-ph/9810535].
1952: %%CITATION = HEP-PH 9810535;%%
1953: 
1954: \bibitem{RSI}
1955: L.~Randall and R.~Sundrum,
1956: %``A large mass hierarchy from a small extra dimension,''
1957: Phys.\ Rev.\ Lett.\  {\bf 83}, 3370 (1999)
1958: [hep-ph/9905221].
1959: %%CITATION = HEP-PH 9905221;%%
1960: 
1961: \bibitem{RSII}
1962: L.~Randall and R.~Sundrum,
1963: %``An alternative to compactification,''
1964: Phys.\ Rev.\ Lett.\  {\bf 83}, 4690 (1999)
1965: [hep-th/9906064].
1966: %%CITATION = HEP-TH 9906064;%%
1967: 
1968: \bibitem{Lorentz}
1969: C.~Csaki, J.~Erlich and C.~Grojean,
1970: %``Gravitational Lorentz violations and adjustment of the cosmological  constant in asymmetrically warped spacetimes,''
1971: Nucl.\ Phys.\ B {\bf 604}, 312 (2001)
1972: [hep-th/0012143];
1973: %%CITATION = HEP-TH 0012143;%%
1974: C.~Csaki, J.~Erlich and C.~Grojean,
1975: %``The cosmological constant problem in brane-worlds and gravitational  Lorentz violations,''
1976: gr-qc/0105114;
1977: %%CITATION = GR-QC 0105114;%%
1978: C.~Grojean, F.~Quevedo, G.~Tasinato and I.~Zavala C.,
1979: %``Branes on charged dilatonic backgrounds: Self-tuning, Lorentz  violations and cosmology,''
1980: JHEP {\bf 0108}, 005 (2001)
1981: [hep-th/0106120].
1982: %%CITATION = HEP-TH 0106120;%%
1983: 
1984: \bibitem{DRV}
1985: A.~C.~Davis, C.~Rhodes and I.~Vernon,
1986: %``Branes on the horizon,''
1987: JHEP {\bf 0111}, 015 (2001)
1988: [hep-ph/0107250].
1989: %%CITATION = HEP-PH 0107250;%%
1990: 
1991: \bibitem{LVBounds}
1992: S.~K.~Lamoreaux, J.~P.~Jacobs, B.~R.~Heckel, F.~J.~Raab and E.~N.~Fortson,
1993: %``New Limits On Spatial Anisotrophy From Optically Pumped He-201 And Hg-199,''
1994: Phys.\ Rev.\ Lett.\  {\bf 57}, 3125 (1986);
1995: %%CITATION = PRLTA,57,3125;%%
1996: R.~S.~Van Dyck, P.~B.~Schwinberg and H.~G.~Dehmelt,
1997: %``New High Precision Comparison Of Electron And Positron G Factors,''
1998: Phys.\ Rev.\ Lett.\  {\bf 59} (1987) 26;
1999: %%CITATION = PRLTA,59,26;%%
2000: S.~M.~Carroll, G.~B.~Field and R.~Jackiw,
2001: %``Limits On A Lorentz And Parity Violating Modification Of Electrodynamics,''
2002: Phys.\ Rev.\ D {\bf 41}, 1231 (1990);
2003: %%CITATION = PHRVA,D41,1231;%%
2004: G.~Gabrielse, D.~Phillips, W.~Quint, H.~Kalinowsky, G.~Rouleau and W.~Jhe,
2005: %``Special Relativity And The Single Anti-Proton: Fortyfold Improved Comparison Of Anti-P And P Charge To Mass Ratios,''
2006: Phys.\ Rev.\ Lett.\  {\bf 74} (1995) 3544;
2007: %%CITATION = PRLTA,74,3544;%%
2008: J.~F.~Wardle, R.~A.~Perley and M.~H.~Cohen,
2009: %``Observational Evidence Against Birefringence over Cosmological Distances,''
2010: Phys.\ Rev.\ Lett.\  {\bf 79}, 1801 (1997)
2011: [astro-ph/9705142];
2012: %%CITATION = HEP-PH 9812418;%%
2013: R.~Bluhm, V.~A.~Kostelecky and N.~Russell,
2014: %``CPT and Lorentz tests in hydrogen and antihydrogen,''
2015: Phys.\ Rev.\ Lett.\  {\bf 82}, 2254 (1999)
2016: [hep-ph/9810269];
2017: %%CITATION = HEP-PH 9810269;%%
2018: V.~A.~Kostelecky and C.~D.~Lane,
2019: %``Constraints on Lorentz violation from clock-comparison experiments,''
2020: Phys.\ Rev.\ D {\bf 60}, 116010 (1999)
2021: [hep-ph/9908504];
2022: %%CITATION = HEP-PH 9908504;%%
2023: V.~A.~Kostelecky,
2024: %``Theory and tests of CPT and Lorentz violation,''
2025: hep-ph/9904467;
2026: %%CITATION = HEP-PH 9904467;%%
2027: R.~Cowsik, G.~Rajalakshmi and B.~V.~Sreekantan,
2028: %``Cosmic Ray Bounds On Violation Of Lorentz Invariance,''
2029: {\it Prepared for 26th International Cosmic Ray Conference (ICRC 99), Salt Lake City, Utah, 17-25 Aug 1999};
2030: D.~Bear, R.~E.~Stoner, R.~L.~Walsworth, V.~A.~Kostelecky and C.~D.~Lane,
2031: %``Limit on Lorentz and CPT violation of the neutron using a two-species  noble-gas maser,''
2032: Phys.\ Rev.\ Lett.\  {\bf 85}, 5038 (2000)
2033: [physics/0007049];
2034: %%CITATION = PHYS-ICS 0007049;%%
2035: D.~F.~Phillips, M.~A.~Humphrey, E.~M.~Mattison, R.~E.~Stoner, R.~F.~Vessot and R.~L.~Walsworth,
2036: %``Limit on Lorentz and CPT violation of the proton using a hydrogen  maser,''
2037: Phys.\ Rev.\ D {\bf 63}, 111101 (2001) [physics/0008230];
2038: %%CITATION = PHYS-ICS 0008230;%%
2039: J.M.~Carmona and J.L.~Cortes, {\it Phys. Lett.} {\bf B494} (2000) 75-80,
2040: hep-th/0012028;
2041: %
2042: T.~Jacobson, S.~Liberati and D.~Mattingly, UMD preprint
2043: hep-ph/0112207.
2044: 
2045: \bibitem{Glashow}
2046: %%CITATION = ASTRO-PH 9705142;%%
2047: S.~R.~Coleman and S.~L.~Glashow,
2048: %``Cosmic ray and neutrino tests of special relativity,''
2049: Phys.\ Lett.\ B {\bf 405}, 249 (1997) [hep-ph/9703240];
2050: %%CITATION = HEP-PH 9703240;%%
2051: %``High-energy tests of Lorentz invariance,''
2052: Phys.\ Rev.\ D {\bf 59}, 116008 (1999) [hep-ph/9812418].
2053: 
2054: 
2055: \bibitem{WillBook}
2056: C.~M.~Will, {\it Theory and Experiment in Gravitational Physics,
2057: 2nd Edition}, (Cambridge University Press, 1993).
2058: 
2059: \bibitem{Will}
2060: For a recent summary of experimental bounds on deviations from General
2061: Relativity, see
2062: C.~M.~Will,
2063: %``The confrontation between general relativity and experiment: A 1998  update,''
2064: gr-qc/9811036;
2065: %%CITATION = GR-QC 9811036;%%
2066: C.~M.~Will,
2067: %``The confrontation between general relativity and experiment,''
2068: Living Rev.\ Rel.\  {\bf 4}, 4 (2001)
2069: [gr-qc/0103036].
2070: %%CITATION = GR-QC 0103036;%%
2071: 
2072: 
2073: 
2074: \bibitem{Taylor}
2075: P.~M.~Mcculloch, J.~H.~Taylor and L.~A.~Fowler,
2076: %``Gravitational Radiation And The Binary Pulsar. (Talk),''
2077: {\it  In *Perth 1979, Proceedings, Gravitational Radiation, Collapsed Objects and Exact Solutions*, 5-11.(See Conference Index)};
2078: J.H. Taylor and P.M. McCulloch, {\it Ann. N.Y. Acad. Sci.}
2079: (1980) 442;
2080: J.H. Taylor and J.M. Weisberg, {\it Ap. J.} (1982)
2081: 908.
2082: 
2083: \bibitem{Carlip}
2084: S.~Carlip,
2085: %``Aberration and the Speed of Gravity,''
2086: Phys.\ Lett.\ A {\bf 267}, 81 (2000)
2087: [gr-qc/9909087].
2088: %%CITATION = GR-QC 9909087;%%
2089: 
2090: \bibitem{protvsphot}
2091: M.~Nagano and A.~A.~Watson,
2092: %``Observations And Implications Of The Ultrahigh-Energy Cosmic Rays,''
2093: Rev.\ Mod.\ Phys.\  {\bf 72}, 689 (2000);
2094: %%CITATION = RMPHA,72,689;%%
2095: F.~Halzen, R.~A.~Vazquez, T.~Stanev and H.~P.~Vankov,
2096: %``The highest energy cosmic ray,''
2097: Astropart.\ Phys.\  {\bf 3}, 151 (1995).
2098: %%CITATION = APHYE,3,151;%%
2099: 
2100: \bibitem{MN}
2101: G.~D.~Moore and A.~E.~Nelson,
2102: %``Lower bound on the propagation speed of gravity from gravitational  Cherenkov radiation,''
2103: JHEP {\bf 0109}, 023 (2001)
2104: [hep-ph/0106220].
2105: %%CITATION = HEP-PH 0106220;%%
2106: 
2107: \bibitem{Shortcuts}
2108: R.~R.~Caldwell and D.~Langlois,
2109: %``Shortcuts in the fifth dimension,''
2110: Phys.\ Lett.\ B {\bf 511}, 129 (2001)
2111: [gr-qc/0103070];
2112: %%CITATION = GR-QC 0103070;%%
2113: E.~Abdalla, B.~Cuadros-Melgar, S.~S.~Feng and B.~Wang,
2114: %``The shortest cut in brane cosmology,''
2115: hep-th/0109024.
2116: %%CITATION = HEP-TH 0109024;%%
2117: 
2118: \bibitem{CosmoShort}
2119: J.~W.~Moffat,
2120: %``Superluminary universe: A Possible solution to the initial value problem in cosmology,''
2121: Int.\ J.\ Mod.\ Phys.\ D {\bf 2}, 351 (1993)
2122: [gr-qc/9211020];
2123: %%CITATION = GR-QC 9211020;%%
2124: M.~A.~Clayton and J.~W.~Moffat,
2125: %``A scalar-tensor cosmological model with dynamical light velocity,''
2126: Phys.\ Lett.\ B {\bf 506}, 177 (2001)
2127: [gr-qc/0101126];
2128: %%CITATION = GR-QC 0101126;%%
2129: D.~J.~Chung and K.~Freese,
2130: %``Can geodesics in extra dimensions solve the cosmological horizon  problem?,''
2131: Phys.\ Rev.\ D {\bf 62}, 063513 (2000)
2132: [hep-ph/9910235];
2133: %%CITATION = HEP-PH 9910235;%%
2134: %``Cosmological challenges in theories with extra dimensions and remarks  on the horizon problem,''
2135: Phys.\ Rev.\ D {\bf 61}, 023511 (2000)
2136: [hep-ph/9906542];
2137: %%CITATION = HEP-PH 9906542;%%
2138: A.~Albrecht and J.~Magueijo,
2139: %``A time varying speed of light as a solution to cosmological puzzles,''
2140: Phys.\ Rev.\ D {\bf 59}, 043516 (1999)
2141: [astro-ph/9811018];
2142: %%CITATION = ASTRO-PH 9811018;%%
2143: P.~P.~Avelino and C.~J.~Martins,
2144: %``Does a varying speed of light solve the cosmological problems?,''
2145: Phys.\ Lett.\ B {\bf 459}, 468 (1999)
2146: [astro-ph/9906117];
2147: %%CITATION = ASTRO-PH 9906117;%%
2148: S.~Liberati, B.~A.~Bassett, C.~Molina-Paris and M.~Visser,
2149: %``Chi Variable-Speed-Of-Light Cosmologies,''
2150: Nucl.\ Phys.\ Proc.\ Suppl.\  {\bf 88}, 259 (2000)
2151: [astro-ph/0001481];
2152: %%CITATION = ASTRO-PH 0001481;%%
2153: J.~D.~Barrow and J.~Magueijo,
2154: %``Solving the flatness and quasi-flatness problems in Brans-Dicke  cosmologies with a varying light speed,''
2155: Class.\ Quant.\ Grav.\  {\bf 16}, 1435 (1999)
2156: [astro-ph/9901049].
2157: %%CITATION = ASTRO-PH 9901049;%%
2158: 
2159: \bibitem{GravCosm}
2160: S. Weinberg, {\it Gravitation and Cosmology}, John Wiley \& Sons,
2161: 1972.
2162: 
2163: \bibitem{Nordtvedt}
2164: K. Nordtvedt, Astrophys. J. 320 (1987) 871.
2165: 
2166: %\cite{BL}
2167: \bibitem{BL}
2168: C.~P.~Burgess and D.~London,
2169: Phys.\ Rev.\ Lett.\ {\bf 69} (1992) 3428; Phys.\ Rev.\ D {\bf 48}
2170: (1993) 4337 [hep-ph/9203216].
2171: %On Anomalous Gauge Boson Couplings and Loop Calculations
2172: %%CITATION = PRLTA,69,3428;%%
2173: %``Uses and abuses of effective Lagrangians,''
2174: %%CITATION = HEP-PH 9203216;%%
2175: 
2176: \bibitem{GRW}
2177: G.~F.~Giudice, R.~Rattazzi and J.~D.~Wells,
2178: %``Quantum gravity and extra dimensions at high-energy colliders,''
2179: Nucl.\ Phys.\ B {\bf 544}, 3 (1999)
2180: [hep-ph/9811291].
2181: %%CITATION = HEP-PH 9811291;%%
2182: 
2183: \bibitem{HLZ}
2184: T.~Han, J.~D.~Lykken and R.~J.~Zhang,
2185: %``On Kaluza-Klein states from large extra dimensions,''
2186: Phys.\ Rev.\ D {\bf 59}, 105006 (1999)
2187: [hep-ph/9811350].
2188: %%CITATION = HEP-PH 9811350;%%
2189: 
2190: \bibitem%[adelberger]
2191: {adelberger}
2192: {C.~D.~Hoyle, U.~Schmidt, B.~R.~Heckel,
2193: E.~G.~Adelberger, J.~H.~Gundlach, D.~J.~Kapner and H.~E.~Swanson,
2194: Phys.\ Rev.\ Lett.\  {\bf 86}, 1418 (2001) [hep-ph/0011014].
2195: }
2196: %%CITATION = HEP-PH 0011014;%%
2197: 
2198: \bibitem{ADDBounds}
2199: S.~Cullen and M.~Perelstein,
2200: %``SN1987A constraints on large compact dimensions,''
2201: Phys.\ Rev.\ Lett.\  {\bf 83}, 268 (1999) [arXiv:hep-ph/9903422];
2202: %%CITATION = HEP-PH 9903422;%%
2203: V.~D.~Barger, T.~Han, C.~Kao and R.~J.~Zhang,
2204: %``Astrophysical constraints on large extra dimensions,''
2205: Phys.\ Lett.\ B {\bf 461}, 34 (1999) [arXiv:hep-ph/9905474];
2206: %%CITATION = HEP-PH 9905474;%%
2207: C.~Hanhart, J.~A.~Pons, D.~R.~Phillips and S.~Reddy,
2208: %``The likelihood of GODs' existence: Improving the SN 1987a constraint on
2209: the size of large compact dimensions,'' Phys.\ Lett.\ B {\bf 509},
2210: 1 (2001) [arXiv:astro-ph/0102063];
2211: %%CITATION = ASTRO-PH 0102063;%%
2212: S.~Hannestad and G.~Raffelt,
2213: %``New supernova limit on large extra dimensions,''
2214: Phys.\ Rev.\ Lett.\  {\bf 87}, 051301 (2001)
2215: [arXiv:hep-ph/0103201].
2216: %%CITATION = HEP-PH 0103201;%%
2217: 
2218: \bibitem{ADDGBounds}
2219: L.~J.~Hall and D.~R.~Smith,
2220: %``Cosmological constraints on theories with large extra dimensions,''
2221: Phys.\ Rev.\ D {\bf 60}, 085008 (1999) [arXiv:hep-ph/9904267];
2222: %%CITATION = HEP-PH 9904267;%%
2223: N.~Arkani-Hamed, S.~Dimopoulos and G.~R.~Dvali,
2224: %``Phenomenology, astrophysics and cosmology of theories with
2225: sub-millimeter dimensions and TeV scale quantum gravity,'' Phys.\
2226: Rev.\ D {\bf 59}, 086004 (1999) [arXiv:hep-ph/9807344].
2227: %%CITATION = HEP-PH 9807344;%%
2228: 
2229: \bibitem{LED} D.~Atwood, C.~P.~Burgess, E.~Filotas, F.~Leblond,
2230: D.~London and I.~Maksymyk,
2231: %``Supersymmetric large extra dimensions are small and/or numerous,''
2232: Phys.\ Rev.\ D {\bf 63}, 025007 (2001) [arXiv:hep-ph/0007178].
2233: %%CITATION = HEP-PH 0007178;%%
2234: 
2235: \bibitem{Ellis}
2236: J.~R.~Ellis, K.~Farakos, N.~E.~Mavromatos, V.~A.~Mitsou and D.~V.~Nanopoulos,
2237: %``Astrophysical probes of the constancy of the velocity of light,''
2238: Astrophys.\ J.\  {\bf 535}, 139 (2000)
2239: [astro-ph/9907340].
2240: %%CITATION = ASTRO-PH 9907340;%%
2241: 
2242: \end{thebibliography}
2243: 
2244: \end{document}
2245: