1: \documentclass{JHEP3} % 10pt is ignored!
2: \usepackage{epsfig}
3: \epsfclipon
4:
5: %def del \ssubsubsection
6: \def\ssubsubsection#1{\vspace{3mm} \noindent \textbf{#1} \\ \vspace{-3mm} \\ \noindent}
7:
8: % Jim's macros
9: \newcommand{\sperp}{{\scriptscriptstyle\perp}}
10: \newcommand{\spm}{{\scriptscriptstyle\pm}}
11: \newcommand{\nco}{\newcommand}
12: \nco{\beq}{\begin{equation}} \nco{\eeq}{\end{equation}}
13: \nco{\beqa}{\begin{eqnarray}} \nco{\eeqa}{\end{eqnarray}}
14: \nco{\lra}{\leftrightarrow}
15: \def\sfrac#1#2{{\textstyle{#1\over #2}}}
16: \def\eps{\epsilon}
17: \def\sgn{{\rm sgn}}
18: \nco{\sss}{\scriptscriptstyle} \nco{\dphi}{\varphi}
19: \nco{\lsim}{\mbox{\raisebox{-.6ex}{~$\stackrel{<}{\sim}$~}}}
20: \nco{\gsim}{\mbox{\raisebox{-.6ex}{~$\stackrel{>}{\sim}$~}}}
21: \def\VEV#1{{\langle #1 \rangle}}
22: \def\wt{\widetilde}
23: \def\pref#1{(\ref{#1})}
24:
25: \def\Hp{H_+}
26: \def\hp{h_+}
27: \def\Hm{H_-}
28: \def\hm{h_-}
29:
30: \def\Hbr{{\overline{H}}}
31: \def\hbr{{\overline{h}}}
32: \def\kbr{{\overline{k}}}
33:
34: \title{\Large Effective Field Theories and Inflation}
35:
36: \author{C.P.\ Burgess, J.M. Cline\\
37: Physics Department, McGill University,
38: 3600 University Street, Montr\'eal, Qu\'ebec, Canada H3A 2T8\\
39: E-mail: \email{cliff@physics.mcgill.ca},
40: \email{jcline@physics.mcgill.ca}, }
41: \author{and R.~ Holman\\ Physics Department, Carnegie Mellon
42: University, Pittsburgh PA 15213\\
43: E-mail: \email{rh4a@andrew.cmu.edu}}
44:
45: \preprint{McGill-03/12}
46:
47: \keywords{Cosmology; Inflation}
48:
49: %\received{\today} %%
50: %\accepted{\today} %% These are for published papers.
51: %\JHEP{12(2001)999} %%
52:
53: \abstract{We investigate the possible influence of
54: very-high-energy physics on inflationary predictions focussing on
55: whether effective field theories can allow effects which are
56: parametrically larger than order $H^2/M^2$, where $M$ is the scale
57: of heavy physics and $H$ is the Hubble scale at horizon exit. By
58: investigating supersymmetric hybrid inflation models, we show that
59: decoupling does not preclude heavy-physics having effects for the
60: CMB with observable size even if $H^2/M^2 \ll O(1\%)$, although
61: their presence can only be inferred from observations given some
62: {\it a priori} assumptions about the inflationary mechanism. Our
63: analysis differs from the results of hep-th/0210233, in which
64: other kinds of heavy-physics effects were found which could alter
65: inflationary predictions for CMB fluctuations, inasmuch as the
66: heavy-physics \emph{can} be integrated out here to produce an
67: effective field theory description of low-energy physics. We
68: argue, as in hep-th/0210233, that the potential presence of
69: heavy-physics effects in the CMB does {\it not} alter the
70: predictions of inflation for generic models, but {\it does} make
71: the search for deviations from standard predictions worthwhile. }
72:
73: \begin{document}
74:
75: \section{Introduction and Discussion \label{section:intro}}
76: %
77: The recent inflationary literature contains considerable
78: discussion of whether or not detailed observations of fluctuations
79: in the temperature of the Cosmic Microwave Background (CMB) can be
80: used to infer the properties of extremely high energies ---
81: usually assumed to be above the Planck scale. This discussion is
82: particularly timely due to the arrival of ever-more-accurate
83: measurements of these fluctuations \cite{boomerang,dasi}, most
84: recently by the Wilkinson Microwave Anisotropy Probe (WMAP)
85: collaboration \cite{WMAP}. There is likely to be even further
86: improvement in the not-too-distant future \cite{Planck}.
87:
88: Broadly speaking, two points of view have emerged from this
89: discussion.
90: %
91: \begin{enumerate}
92: \item CMB fluctuations can depend on the details of high-energy
93: (trans-Planckian) physics, and so represents an opportunity to
94: probe these otherwise inaccessible scales \cite{tp1,tp2,tp3}.
95: %
96: \item General decoupling arguments require the influence of physics
97: at scale $M \gg H$ to contribute at most of order $H^2/M^2$ to
98: observable late-time effects, where $H$ is the Hubble scale at
99: horizon exit. This precludes the intrusion of higher-energy
100: physics into CMB fluctuations \cite{shenker}.
101: %
102: \end{enumerate}
103:
104: Clearly much is at stake. On the one hand, given specific guesses
105: for what trans-Planckian physics might be, observable effects for
106: the CMB have been calculated. Although some of these calculations
107: remain controversial \cite{dSinconsistent} -- in particular the
108: choice of nonstandard vacua in de Sitter space -- it is hard to
109: argue trans-Planckian physics cannot interfere with inflationary
110: predictions without better understanding what trans-Planckian
111: physics is.
112:
113: On the other hand, if general decoupling arguments do not apply in
114: an inflationary context then the very {\it predictability} of
115: inflationary models is lost. Indeed, since the nature of the
116: higher-energy physics is at present unknown any of its
117: implications must be included under the general heading of
118: theoretical uncertainty when comparing with experiments. In this
119: sense decoupling is a prerequisite for using the properties of the
120: CMB as evidence for an earlier inflationary period
121: \cite{WMAPinflation} in the first place.
122:
123: To the extent that the central issue is the validity and
124: implications of decoupling during inflation, it may be addressed
125: without invoking unknown trans-Planckian physics. The conclusions
126: drawn might then also be applicable to trans-Planckian physics to
127: the extent that it also satisfies the assumptions of the analysis.
128: Here (and in \cite{us}) we use simple sub-Planckian field theories
129: to explore these issues. Within this context two separate
130: questions may be addressed.
131: %
132: \begin{enumerate}
133: %
134: \item Is it required that higher-energy physics decouple at all,
135: in the sense that its low-energy effects must be described in
136: terms of a low-energy effective field theory?
137: %
138: \item Given that the low-energy implications of heavy physics
139: {\it can} be described by an effective field theory during the
140: epoch of horizon exit, need its influence for the CMB be limited
141: to effects which are of order $H^2/M^2$?
142: %
143: \end{enumerate}
144: %
145: Ref.~\cite{us} addresses the first of these questions, and shows
146: that oscillating background fields before horizon exit can
147: invalidate an effective-field-theory description. They can do so
148: by preventing the time evolution of the relevant modes of the
149: inflaton from being adiabatic, which is also an obstacle to using
150: an effective-lagrangian description outside of a cosmological
151: context.
152:
153: In this paper we address the second question of whether an
154: effective field theory can produce detectable effects even if
155: those of order $H^2/M^2$ are too small to be observable. We here
156: show that for some models heavy physics generates contributions to
157: the inflaton potential which depend logarithmically on the heavy
158: mass $M$, and so contribute to the slow-roll parameters an amount
159: of order $M_0^2/M^2$, with $M_0 \gg H$. In so doing we also show
160: that since the CMB measurements in themselves only probe the
161: inflaton potential in a limited way, at present the influence of
162: heavy physics can only be inferred using CMB measurements given
163: some sort of {\it a priori} assumptions about the nature of the
164: physics which is responsible for making the inflaton potential
165: flat in the first place.
166:
167: Although the thrust of both of these conclusions is that
168: high-energy physics {\it can} intrude into CMB fluctuations, we
169: argue that they do so in a way which does {\it not} introduce
170: uncontrollable theoretical errors into the predictions of
171: inflationary models, and so do not undermine the successful
172: comparison of these predictions with observations. They do not do
173: so because our basic conclusion is that the criteria for
174: decoupling in inflation are precisely the same criteria which
175: apply in other non-cosmological contexts. In particular, the vast
176: majority of high-energy effects {\it do} decouple and so cannot
177: alter standard predictions. But just as in other areas of physics,
178: some specific kinds of effective interactions can be sensitive to
179: higher-energy details and these are worth scrutinizing for the
180: information they may contain about the microscopic physics of very
181: small distances. In this sense our results, like those of
182: \cite{us}, represent in some ways the best of all possible worlds.
183:
184: Our presentation is organized as follows. Section \ref{sec:toy}
185: starts with a toy model of hybrid inflation consisting of two
186: scalar fields having very different masses. We use this model to
187: explicitly integrate out the heavier of the two (the heavy
188: physics) at one loop to see its effects on the lighter scalar (the
189: inflaton). This example illustrates first that, in the generic
190: case heavy-field loop contributions to the inflaton potential are
191: large and dangerous, since they tend to ruin inflation by
192: destroying the flatness of the inflaton potential. This is one of
193: the well-known naturalness problems of inflation, and any careful
194: treatment of the effects for inflation of higher-energy physics
195: must properly address this.
196:
197: We can also use this example to show that even if the above
198: naturalness problems are addressed, there is a further obstruction
199: to identifying heavy-physics effects within the observed
200: fluctuations of the CMB if the inflaton is really rolling slowly
201: at the epoch of horizon exit. This is because the inflationary
202: predictions in this case only sample the first few derivatives of
203: the inflaton potential, and these can generically be adjusted by
204: small changes in the renormalizable inflaton couplings. Because of
205: this a detection of heavy-physics effects in inflation is only
206: possible within the context of a specific inflationary model, for
207: which the inflaton couplings can be subject to {\it a priori}
208: conditions (such as those due to symmetries).
209:
210: Section \ref{sec:flat} sharpens the analysis by addressing these
211: two issues within a supersymmetric model. Supersymmetry provides
212: an attractive framework for addressing the above questions because
213: supersymmetry both controls the naturalness issue and provides
214: {\it a priori} constraints on the form of the low-energy inflaton
215: potential. Section \ref{subsec:global} examines these issues
216: within a standard globally supersymmetric hybrid-inflation model.
217: The light scalar for this model parameterizes a precisely flat
218: direction of the classical potential, whose degeneracy is lifted
219: purely by the virtual effects of heavy fields. In this model it is
220: therefore purely heavy-physics effects which make the difference
221: between the success and failure of the theory as a model of
222: inflation, since it is their small size which explains why the
223: inflaton potential is very shallow (but not exactly flat). Simple
224: modifications of this higher-energy physics produce sizeable
225: changes in the predictions for the CMB precisely because in these
226: models it is the higher-energy physics which controls the entire
227: effect.
228:
229: Section \ref{subsec:sugra} generalizes the model of section
230: \ref{subsec:global} to supergravity. The main message of this
231: section is that such an extension is possible despite the
232: well-known $\eta$ problem of supergravity models. We illustrate
233: this using the example of D-term inflation.
234:
235: We emphasize that all of the examples we use are extremely
236: orthodox, and have been considered in various contexts in the
237: literature. Our purpose in bringing them all together here is to
238: first show that existing models already provide explicit examples of how
239: higher-energy physics can have important effects during inflation,
240: and so to illustrate that such effects can happen in ordinary
241: theories which decouple, without relying on other exotic
242: properties (like violations of Lorentz invariance, for example).
243: A secondary goal is to raise the bar for putative models of
244: trans-Planckian physics, by arguing that any serious candidate for
245: this must: ($i$) establish that the proposed trans-Planckian
246: physics does decouple from lower-energy phenomena so as to not
247: completely sacrifice the predictivity of lower-energy physics, and
248: ($ii$) contain a precise model of the low-energy inflaton dynamics
249: as a benchmark against which the high-energy physics effects can
250: be compared.
251:
252: \section{Inflation in a Toy Effective Theory}\label{sec:toy}
253: %
254: In this section we accomplish two things using a simple toy model
255: involving an inflaton $\phi$ and some heavy fields $\chi_i$.
256: First, we explicitly integrate out the heavy fields to derive the
257: dominant terms in the resulting low-energy effective field theory
258: for $\phi$. Then we apply this result to an inflationary model and
259: show how these heavy-field contributions are poison to
260: inflationary models, since they destroy the flatness of the
261: inflaton potential. This well-known naturalness problem shows that
262: the virtual effects of high-energy fields for inflation are
263: generically too {\it large} rather than too small. This
264: calculation is standard \cite{RobertRMP} and is mainly used to
265: motivate the calculations of the next section, where we repeat the
266: analysis using supersymmetric examples. A reader familiar with
267: these issues, and who is in a hurry, can skip directly to section
268: \ref{sec:flat}.
269:
270: \subsection{The Model}\label{subsec:model}
271: %
272: Consider the following toy model of very-high-energy physics,
273: which is a very minor generalization of a standard
274: hybrid-inflation model \cite{hybrid}:
275: %
276: \beqa \label{eq:modeldef}
277: - \, {{\cal L} \over \sqrt{-g}} &=& \sfrac12 \, \partial_\mu \phi \,
278: \partial^\mu \phi + \sfrac12 \, \partial_\mu \chi_i \, \partial^\mu \chi_i +
279: V(\phi,\chi) , \\
280: \hbox{with} \qquad V(\phi,\chi) &=& V_{\rm inf}(\phi) - \sfrac12 \,
281: m^2 \chi_i^2 + \sfrac12 \, g \, \chi_i^2 \phi^2 . \nonumber
282: \eeqa
283: %
284: We assume the $N$ massive scalars,\footnote{We introduce $N$ heavy
285: fields simply to amplify the effects of the $\chi$ fields, and
286: much of our discussion applies equally well if $N=1$. Perturbative
287: quartic self-interactions for $\chi$ can also be included without
288: materially affecting our discussion.} $\chi_i, i=1,...,N$
289: --- which represent the heavy physics whose lower-energy influence
290: we wish to determine --- satisfy $\langle \chi_i \rangle = 0$,
291: thereby excluding the non-adiabatic effects discussed in
292: ref.~\cite{us}. For the simplicity of later formulae we choose the
293: couplings to be invariant under the $O(N)$ which rotates the
294: $\chi_i$'s amongst themselves. In eq.~\pref{eq:modeldef} $\phi$ is
295: the inflaton, whose potential
296: %
297: \beq \label{eq:Vinf}
298: V_{\rm inf}(\phi) = \rho + \sfrac12 \, m_\phi^2 \, \phi^2 +
299: \sfrac14 \, \lambda \, \phi^4
300: \eeq
301: %
302: is chosen to ensure sufficient $e$-foldings of expansion before an
303: eventual inflationary exit and reheating. This is ensured if the
304: two slow-roll parameters \cite{liddlelyth} satisfy\footnote{We use
305: here the rationalized Planck mass, $M_p^{-2} = 8 \pi G$.}
306: %
307: \beqa \label{eq:slowroll}
308: \epsilon_0(\phi) &=& \frac12 \, \left( {M_p \, V'_{\rm inf} \over
309: V_{\rm inf}} \right)^2 = \left( {M_p \, \phi \, (m_\phi^2 +
310: \lambda \, \phi^2) \over \rho + \sfrac12 \, m_\phi^2 \, \phi^2 +
311: \sfrac14 \, \lambda \, \phi^4 } \right)^2 \ll 1 \nonumber \\
312: \eta_0(\phi) &=& \left( {M_p^2 \, V_{\rm inf}'' \over
313: V_{\rm inf}} \right) = \left( {M_p^2 \, (m_\phi^2 + 3
314: \lambda \, \phi^2) \over \rho + \sfrac12 \, m_\phi^2 \, \phi^2 +
315: \sfrac14 \, \lambda \, \phi^4 } \right) \ll 1 \, .
316: \eeqa
317: %
318: For instance, these conditions are satisfied if $m_\phi^2 M_p^2
319: \ll \rho$ and if we choose initial conditions for which $\chi_i =
320: 0$ and $\rho/m_\phi^2 \gg \phi^2 \gg m^2/g$ (provided $\lambda$ is
321: small enough also to ensure $\lambda \phi^2 \ll m_\phi^2$
322: throughout this range of $\phi$). Inflation then occurs with $H^2
323: = V_{\rm inf}(\phi)/(3 M_p^2) \approx \rho/(3 M_p^2)$ while $\phi$
324: rolls slowly down to $\phi_{\rm end}$, where inflation ends.
325: Inflation ends either when the slow-roll parameters become of
326: order unity, or when $\phi^2 \sim m^2/g$, at which point $\chi_i$
327: moves quickly away from zero, and $\phi_{\rm end}$ is the field
328: determined by whichever of these occurs first.
329:
330: In order to study the effects of high energy physics we assume the
331: masses,
332: %
333: \beq
334: \label{eq:Mphi}
335: M^2(\phi) = -m^2 + g \, \phi^2,
336: \eeq
337: %
338: to be much larger than the Hubble scale, $H^2$, during horizon
339: exit. In order to keep the analysis under control we also assume
340: the $\chi_i$ fields to be sub-Planckian throughout the
341: inflationary slow roll of $\phi$: $M^2(\phi) \ll M_p^2$.
342:
343: \subsection{Integrating out the Heavy Fields}\label{subsec:heavy}
344: %
345: Under the above assumptions the heavy fields may be explicitly
346: integrated out and an effective-field-theory analysis applies.
347: Since our main interest is in effects which do not decouple we
348: focus on those terms in the effective theory which are
349: unsuppressed by powers of $1/M$. These come in two types: ($i$)
350: relevant interactions, which are proportional to positive powers
351: of $M$; or ($ii$) marginal interactions, which grow
352: logarithmically with $M$.
353:
354: A simple calculation of the virtual effects of the heavy scalars
355: on the light scalar potential is obtained by matching the one-loop
356: corrected effective potential for the full theory with the
357: one-loop effective potential in the theory involving only $\phi$,
358: but including effective interactions. This gives the following
359: result:
360: %
361: \beqa \label{eq:effpot}
362: V_{\rm eff}(\phi) &=& V_{\rm inf}(\phi) + \Delta V(\phi), \\
363: \hbox{with} \qquad \Delta V(\phi) &=& V_{\rm ct}(\phi) + \frac{N}{64 \pi^2}
364: M^4(\phi) \, \ln \left( {M^2(\phi) \over \mu^2} \right) \, ,
365: \nonumber
366: \eeqa
367: %
368: where $V_{\rm inf}$ now denotes the renormalized inflaton
369: potential, eq.~\pref{eq:Vinf}, with constants $\rho(\mu)$,
370: $m_\phi^2(\mu)$ and $\lambda(\mu)$ chosen according to a
371: renormalization prescription which is described in detail below.
372: $V_{\rm ct}$ contains the counter-terms, $\delta \rho$, $\delta
373: m_\phi^2$ and $\delta \lambda$, which enforce the renormalization
374: condition, and $\mu$ is a floating scale which also depends on the
375: renormalization scheme. As usual, the implicit dependence of
376: $V_{\rm inf}$ on $\mu$ is just such as to ensure that $\mu$
377: cancels in physical observables.
378:
379: For inflationary purposes it is convenient to cast the
380: renormalization conditions in terms of the slow-roll parameters
381: since we must really demand that $V_{\rm eff}$, rather then
382: $V_{\rm inf}$, produces sufficient inflation. For our laterpurposes, a convenient way to do so is to require that $V_{\rm
383: eff}$ and $V_{\rm inf}$ share the same value for the Hubble
384: constant and the two slow-roll parameters at a particular point in
385: field space, which we can choose to be the point when a specific
386: mode $k_*$ leaves the horizon. We denote the time of horizon exit
387: for this mode by $t_*$, where $H(t_*) = k_{*{\rm phys}} =
388: k_*/a(t_*)$, and we further denote the value of the inflaton field
389: at this time by $\phi(t_*) = \phi_*$. Then this renormalization
390: condition states that $V_{\rm eff}(\phi_*)$, $\epsilon(\phi_*)$
391: and $\eta(\phi_*)$ are given by their tree-level expressions in
392: terms of $\rho$, $m_\phi$ and $\lambda$:
393: %
394: \beqa \label{eq:rencond}
395: V_* &=& V_{\rm eff}(\phi_*) = \rho + \sfrac12 \, m_\phi \, \phi_*^2 +
396: \sfrac14 \, \lambda \, \phi_*^4 = 3 \, M_p^2 \, H_*^2\nonumber \\
397: \epsilon_* &=& \epsilon_0(\phi_*) = \frac12 \, \left[ {M_p \, \phi_* \,
398: ( m_\phi^2 + \lambda \, \phi_*^2) \over V_{\rm eff}(\phi_*)} \right]^2
399: = \frac12 \, \left[ {\phi_* \,( m_\phi^2
400: + \lambda \, \phi_*^2) \over 3 \, M_p \, H_*^2} \right]^2 \\
401: \eta_* &=& \eta_0(\phi_*) = {M_p^2 \, (m_\phi^2 + 3 \, \lambda \, \phi_*^2)
402: \over V_{\rm eff}(\phi_*)} = {(m_\phi^2 + 3 \, \lambda \, \phi_*^2)
403: \over 3 \, H_*^2} \, . \nonumber
404: \eeqa
405: %
406: The subscript `0' on the quantities $\epsilon_0(\phi)$ and
407: $\eta_0(\phi)$ indicates that their functional dependence is as
408: calculated using the potential $V_{\rm inf}$, as in
409: eq.~\pref{eq:slowroll}.
410:
411: These conditions amount to the following requirements for $\Delta
412: V$: $\Delta V(\phi_*) = \Delta V'(\phi_*) = \Delta V''(\phi_*) =
413: 0$, and so
414: %
415: \beq \label{eq:DeltaV}
416: \Delta V(\phi) = {N \over 64 \pi^2} \left\{ M^4(\phi) \left[
417: \ln \left( {M^2(\phi) \over M^2_*} \right) - \sfrac32 \right]
418: + \sfrac12 \, M_*^2 \Bigl[4 \, M^2(\phi) - M_*^2 \Bigr]
419: \right\},
420: \eeq
421: %
422: where $M_* = M(\phi_*)$.
423:
424: Using the following derivatives:
425: %
426: \beqa \label{eq:dVeff}
427: \Delta V'(\phi) &=& {g \, N \, \phi \over 16 \pi^2}
428: \left[ (m^2 + g \, \phi^2) \ln \left( {M^2 \over M_*^2} \right) -
429: g ( \phi^2 - \phi_*^2 ) \right] ,\nonumber \\
430: \Delta V''(\phi) &=& { g \, N \over 16 \pi^2 } \left[ (m^2 + 3 g \,
431: \phi^2) \ln \left( {M^2 \over M_*^2} \right) - g( \phi^2 -
432: \phi_*^2) \right] , \nonumber \\
433: \Delta V'''(\phi) &=& {g^2 \, N \, \phi \over 8 \pi^2} \left[
434: 3 \ln \left( { M^2 \over M_*^2} \right) + {2 g \,
435: \phi^2 \over M^2} \right],
436: \eeqa
437: %
438: and taking $M^2 \approx g \phi^2 \gg m^2$ and $\Delta V \ll V_{\rm
439: inf} \approx \rho$, we find $H^2 \approx \rho/(3 M_p^2)$ and the
440: following expressions for the slow-roll parameters:
441: %
442: \beqa \label{eq:dslowroll}
443: (2 \, \epsilon)^{1/2} &\approx& (2 \, \epsilon_0)^{1/2} + {g^2 \, N \, \phi \over
444: 48 \pi^2 \, M_p \, H^2} \left[ \phi^2 \ln \left( {\phi^2 \over \phi_*^2} \right) -
445: ( \phi^2 - \phi_*^2 ) \right] \nonumber \\
446: \eta &\approx& \eta_0 + { g^2 \, N \over 48 \pi^2 \, H^2} \left[ 3\,
447: \phi^2 \ln \left( {\phi^2 \over \phi_*^2} \right) - ( \phi^2 -
448: \phi_*^2) \right] \, ,
449: \eeqa
450: %
451: as well as the second-order slow-roll quantity:
452: %
453: \beq \label{eq:xidef}
454: \xi^2 = (2 \, \epsilon)^{1/2} \, \left( {M_p^3 V''' \over V} \right) \approx
455: (2 \, \epsilon)^{1/2} \, {M_p \, \phi \over H^2} \,
456: \left\{ 2 \lambda + \, {g^2 \, N \over 24 \pi^2 } \left[
457: 3 \ln \left( { \phi^2 \over \phi_*^2} \right) + 2 \right] \right\}.
458: \eeq
459:
460: For our purposes, there are two lessons to be learned from these
461: expressions: the naturalness problems they imply, and the
462: obstruction they raise to the inference of heavy-physics
463: properties purely using measurements of the CMB.
464:
465: \ssubsubsection{Naturalness Problems}\label{subsubsec:naturalness}
466: %
467: The biggest difficulty with these expressions is in maintaining
468: inflation itself. Although we've ensured (by construction) that
469: the $\phi$ motion is sufficiently slow near $\phi = \phi_*$, this
470: is not in itself sufficient to obtain the more than 50
471: $e$-foldings of inflation which are required for successful
472: cosmology. In particular, this requires $\eta$ to remain small
473: over a significant range of $\phi$, which requires
474: %
475: \beq \label{eq:dphivsH}
476: \left| \phi^2 - \phi^2_* \right| \ll {48 \pi^2 \,
477: H^2 \over g^2 \, N} \, .
478: \eeq
479: %
480: This is typically difficult to satisfy unless $g$ is extremely
481: small. For instance, if we take $H$ to be approximately constant
482: over the $N_e$ $e$-foldings of inflation between horizon exit and
483: inflation's end, and if we take $dV_{\rm inf}/d\phi \approx
484: m_\phi^2 \, \phi$ over this region, we have $\phi_*^2 \approx
485: (m^2/g) \exp\left[(2 m_\phi^2 \, N_e)/(3 H^2) \right]$. In this
486: case we see eq.~\pref{eq:dphivsH} implies
487: %
488: \beq
489: \left| \exp\left[ \frac{2 m_\phi^2 \, N_e}{3 H^2} \right] - 1
490: \right| \ll \frac{48 \pi^2}{g \, N} \,
491: \left( \frac{H^2}{m^2} \right) \, ,
492: \eeq
493: %
494: and so even if we have already assumed $H^2/m_\phi^2 \sim N_e
495: \gsim 50$, we must now in addition tune $m$ to satisfy $N m^2 \ll
496: 48 \pi^2 H^2/g$.
497:
498: We see that heavy loops can have big effects because they compete
499: with unusually small low-energy interactions. The low-energy
500: interactions are small precisely because inflation requires the
501: low-energy inflaton potential must be chosen to be so very flat.
502: In these circumstance generic kinds of heavy physics not only do
503: not decouple, they can completely ruin inflation.
504:
505: \ssubsubsection{Detecting Heavy Physics Using the CMB}\label{subsubsec:detect}
506: %
507: Eq.~\pref{eq:dslowroll} implies another obstruction to learning
508: about higher-energy physics using inflationary predictions for the
509: CMB, independent of the naturalness issues associated with
510: obtaining the slow roll itself. This obstruction arises because if
511: the inflaton describes a sufficiently slow roll, its effects for
512: the CMB are completely determined by the quantities $H$,
513: $\epsilon$ and $\eta$ evaluated at horizon exit. For instance
514: standard expressions \cite{liddlelyth} for the corrections to the
515: CMB fluctuation spectrum are\footnote{Note that there is
516: a sign error in the formula for $\xi$ in
517: ref.~\cite{liddlelyth}}
518: %
519: \beqa \label{eq:spectrum}
520: \delta_H^2(k_*) &\approx& {V_* \over 150 \pi^2 \, M_p^4 \,
521: \epsilon_*}
522: = {H^2_* \over 50 \pi^2 \, M_p^2 \, \epsilon_*}
523: \nonumber \\
524: n(k_*) &\approx& 1 + 2 \, \eta_* - 6 \, \epsilon_* \\
525: % {\delta_T \over \delta_S} &\approx& \epsilon_* ...\nonumber\\
526: {d n \over d \ln k}(k_*) &\approx& 16 \, \epsilon_* \, \eta_* - 24 \,
527: \epsilon^2_* - 2 \, \xi^2_* . \nonumber
528: \eeqa
529: %
530: Since the above renormalization scheme is designed not to change
531: the slow-roll parameters at horizon exit, it ensures that the
532: heavy fields cannot alter the inflationary predictions for modes
533: near $k = k_*$, to leading order in the slow-roll parameters. To
534: the extent that this is true, and that observations are only
535: sensitive to fluctuation properties near $k_*$, any detection of
536: heavy physics as a distortion of the CMB spectrum must rely on the
537: breakdown of the slow-roll conditions near horizon exit (such as
538: might be true if the preliminary WMAP indications for nonzero
539: $dn/d\ln k$ \cite{WMAPnrun} should prove to be significant).
540:
541: Until more detailed observational information is available, we
542: conclude that the effects of very heavy physics for the CMB can at
543: present only be inferred relative to some {\it a priori}
544: information about the nature of inflationary physics. If, for
545: instance, inflation were believed to be due to a particular
546: mechanism --- such as perhaps the string-motivated brane-inflation
547: proposals of \cite{BI} --- then detailed knowledge of the form for
548: $V_{\rm inf}$ can allow sufficiently large corrections to this
549: form to be inferred from observations.
550:
551: An extreme example of {\it a priori} constraints is the case where
552: $V_{\rm inf}$ is precisely constant, in which case the inflaton
553: potential is {\it entirely} due to loop-generated heavy-physics
554: effects. This situation is actually fairly common for
555: supersymmetric models, which are typically rife with
556: classically-flat directions. In these models the inflaton can be a
557: modulus parameterizing one of these directions, and so the very
558: flatness of the inflaton potential is then partially explained by
559: its origin as a loop-generated effect.
560:
561: \section{Flat Directions and Supersymmetric Models}\label{sec:flat}
562: %
563: In this section we repeat the calculation of the previous section
564: for a supersymmetric example, for which the generic naturalness
565: issues raised above can be controlled. This allows us to more
566: precisely compute the size of high-energy loop effects on CMB
567: fluctuations, and so to illustrate how these effects need not be
568: unobservably small even if their scale is much higher than $H$.
569: They also provide examples wherein the entire inflaton potential
570: arises as such a heavy-physics loop effect, and so for which the
571: very detection of inflationary effects in the CMB is necessarily
572: also a detection of heavy-physics effects. We first describe
573: models in global supersymmetry in section \ref{subsec:global}, for
574: which the calculations are simple. We shall find self-consistency
575: forces us then to generalize these to supergravity, and this is
576: done in section \ref{subsec:sugra}.
577:
578: The results we obtain in this section are simple to state. In the
579: models we consider here, the tree-level inflaton potential is
580: exactly flat, but this flat direction is lifted by virtual loops
581: of heavy particles yielding nonzero slow-roll parameters $\eta$
582: and $\epsilon$. In particular, integrating out the heavy particles
583: produces slow-roll parameters which are suppressed by factors of
584: order $M_0^2/M^2$, where $M$ is the relevant heavy mass scale, and
585: it is this decoupling which makes the slow-roll parameters small.
586: But, and this is the crux of the matter, the reference scale $M_0$
587: is much bigger than $H$, and so these models may be taken as an
588: existence proof that heavy physics can decouple and yet still
589: alter inflationary predictions for the CMB since the figure of
590: merit for deciding the observability of the heavy-physics effects
591: can be larger than $H^2/M^2$.
592:
593: \subsection{Globally Supersymmetric Models}\label{subsec:global}
594: %
595: Part of the attraction of supersymmetric models for hybrid
596: inflation is the ubiquity with which their scalar potentials have
597: flat directions. Two kinds of models have been proposed, which
598: differ according to whether the inflationary potential arises as
599: an $F$-term \cite{Fterm} or a $D$-term \cite{Dterm}. We focus here
600: on $D$-term models in order to avoid the usual $\eta$ problem when
601: we generalize to supergravity.
602:
603: Consider, then, a model containing the chiral multiplets, $\Phi =
604: \{\phi,\psi\}$ and $H_\pm = \{h_\pm,\chi_\pm\}$, coupled to a
605: $U(1)$ gauge multiplet, $V = \{A_\mu,\lambda\}$. We take $\Hp$ and
606: $\Hm$ to carry opposite $U(1)$ charge $\pm e$, and the multiplet
607: $\Phi$ to be neutral. The model's superpotential and K\"ahler
608: potential are
609: %
610: \beq
611: K = \Hp^* \Hp + \Hm^* \Hm + \Phi^* \Phi \qquad \hbox{and}
612: \qquad W = g \Phi \, (\Hp \, \Hm - v^2) \, ,
613: \eeq
614: %
615: where $g$ and $v$ are real constants. The associated scalar
616: potential is $V = V_F + V_D$ where
617: %
618: \beqa
619: V_F &=& g^2 \left( \Bigl| \hp \hm - v^2 \Bigr|^2 + \Bigl| \phi \,
620: \hm \Bigr|^2 + \Bigl| \phi \, \hp \Bigr|^2 \right) \, , \nonumber \\
621: V_D &=& \frac{e^2}{2} \, \Bigl( |\hp|^2 - |\hm|^2 + \xi
622: \Bigr)^2 \, ,
623: \eeqa
624: %
625: where $\xi > 0$ is the Fayet-Iliopoulos term. Notice that $V_D =
626: 0$ implies $|\hm|^2 = |\hp|^2 + \xi$ for any $\phi$ while $V_F =
627: 0$ implies $\phi = 0$ and $\hp \hm = v^2$, so the global minimum
628: is supersymmetric and has
629: %
630: \beq
631: \phi = 0, \qquad |h_\pm|^2 = \frac12 \Bigl( \mp \xi + \sqrt{\xi^2
632: + 4 v^4} \Bigr) \, .
633: \eeq
634:
635: The feature of most interest for the present purposes is the
636: potential's trough at $h_\pm = 0$ and large $|\phi|$, for which
637: $V_{\rm trough}(\phi) = V(h_\pm = 0, \phi) = g^2 \, v^4 + \sfrac12
638: \, e^2 \xi^2$. For sufficiently large $|\phi|$ this is a local
639: minimum in the $h_\pm$ directions, with scalar excitations in
640: these directions having masses
641: %
642: \beq
643: M_\pm^2(\phi) = g^2 |\phi|^2 \pm \sqrt{ g^4 v^4 + e^4 \xi^2} \, .
644: \eeq
645: %
646: Everywhere along this trough the scalar $\phi$ is precisely
647: massless, while linear combinations of the complex scalars $h_\pm$
648: and $h^*_\mp$ are massive, with masses, $M_\pm^2(\phi)$, which are
649: large for large $|\phi|$. Since the $U(1)$ gauge invariance is not
650: broken along the trough, the gauge bosons are massless. The
651: fermion masses along the trough, on the other hand, are zero for
652: the gaugino, $\lambda$, and the chiral-multiplet fermion $\psi$.
653: They are nonzero for the fermions $\chi_\pm$, with mass
654: eigenvalues
655: %
656: \beq
657: m^2_\pm(\phi) = g^2 |\phi|^2 \, .
658: \eeq
659:
660: Along the trough's bottom the tree-level spectrum therefore breaks
661: up into a sector of massless particles,
662: $\{A_\mu,\lambda,\phi,\psi\}$, which do not classically directly
663: couple among themselves, but which do couple to a massive sector,
664: $\{\hp,\chi_+,\hm,\chi_-\}$. In this section we integrate out this
665: massive sector to determine the effective interactions which are
666: generated in this way amongst the light fields, with the goal of
667: using this as a candidate hybrid inflation model \cite{gShybrid}.
668:
669: At the classical level the model does {\it not} describe viable
670: hybrid inflation, with $\phi$ interpreted as the inflaton rolling
671: along the trough's bottom. This is because at the classical level
672: the $\phi$ potential is precisely flat, and so there is nothing to
673: drive the inflaton's slow roll. This is no longer true once loop
674: corrections are included since the virtual heavy fields will
675: generate an inflaton potential.
676:
677: Following the same procedure as for the previous section we obtain
678: the heavy-field contribution to the inflaton potential by matching
679: the one-loop results. It is useful to extend the model to include
680: $N$ charged chiral multiplets, $H^i_\pm$, in an $O(N)$-invariant
681: way, in which case we find the result
682: %
683: \beqa \label{eq:effpotgs}
684: V_{\rm eff}(\phi) &=& \rho + \Delta V(\phi), \\
685: \hbox{with} \qquad \Delta V(\phi) &=& \delta \rho +\frac{2N}{64
686: \pi^2} \sum_{i=\pm} \left[ M_i^4(\phi) \, \ln \left( {M_i^2(\phi)
687: \over \mu^2} \right) - m_i^4(\phi) \, \ln \left( {m_i^2(\phi)
688: \over \mu^2} \right) \right] \, ,
689: \nonumber
690: \eeqa
691: %
692: where $\rho$ is the renormalized (constant) classical potential
693: along the trough (which classically is $\rho = g^2 \, v^4 +
694: \sfrac12 \, e^2 \xi^2$), and $\delta \rho$ is the corresponding
695: counter-term.
696:
697: It is convenient to evaluate this result using the previous
698: expressions $M^2_\pm(\phi) = m^2(\phi) \pm \Delta$ and
699: $m^2_\pm(\phi) = m^2(\phi)$, where $m(\phi) = g|\phi|$ and $\Delta
700: = \sqrt{g^4 v^4 + e^4 \xi^2}$. In terms of these the potential
701: becomes
702: %
703: \beqa \label{eq:effpotgsexpl}
704: \Delta V_{\rm eff}(\phi) = \delta \rho &+& \frac{N}{32
705: \pi^2} \left[m^4(\phi) \, \ln \left( {m^4(\phi) - \Delta^2
706: \over m^4(\phi)} \right) + \Delta^2 \, \ln \left( {m^4(\phi)
707: - \Delta^2 \over \mu^4} \right) \right. \nonumber\\ &+&
708: \left. 2 m^2(\phi) \Delta \, \ln \left(
709: {m^2(\phi) + \Delta
710: \over m^2(\phi) - \Delta} \right) \right] \, ,
711: \nonumber
712: \eeqa
713: %
714: which for $m^2(\phi) \gg \Delta$ becomes
715: %
716: \beq
717: \Delta V_{\rm eff}(\phi) \approx \frac{N \, \Delta^2}{16
718: \pi^2}\left[ \ln \left( {m^2(\phi) \over m_*^2} \right)
719: + {\cal O} \left( {\Delta^2 \over m^4 } \right) \right] \, ,
720: \eeq
721: %
722: where $m_*^2 = m^2(\phi_*) = g^2 |\phi_*|^2$ and we adopt the
723: renormalization condition that $\Delta V$ must vanish when $\phi =
724: \phi_*$.
725:
726: This induced potential causes the inflaton $\varphi = |\phi|$ to
727: roll, and this roll can be slow if $|\phi|$ is sufficiently large,
728: since
729: %
730: \beqa
731: \label{sr}
732: (2 \epsilon)^{1/2} &=& {M_p \over V_{\rm
733: eff}} \left( {\partial V_{\rm eff} \over
734: \partial \varphi} \right) \approx {M_p \,N \, \Delta^2
735: \over 8 \pi^2 \rho \, \varphi}
736: \, , \nonumber \\
737: \eta &=& {M_p^2 \over V_{\rm eff}} \left( {\partial^2
738: V_{\rm eff} \over \partial \varphi^2} \right)
739: \approx - \, {M_p^2 \,N\, \Delta^2 \over 8 \pi^2
740: \rho\, \varphi^2}
741: \, .
742: \eeqa
743:
744: With these expressions we can examine more quantitatively the
745: conditions for inflaton. For simplicity we take for these purposes
746: $e \sim g$ and $\xi \sim v^2$, which implies $\rho \sim g^2 v^4$
747: and $\Delta \sim g^2 v^2$ and so $\Delta^2/\rho \sim g^2$. We also
748: choose $N=1$, although we return to other choices for $N$ at the
749: end. Suppose now $N_e$ $e$-foldings of inflation occur between
750: horizon exit and the end of inflation. During this time the field
751: $\phi$ evolves from $\phi_*$ to $\phi_{\rm end}$, with
752: %
753: \beq
754: \phi_*^2 \approx \phi^2_{\rm end} + \frac{N N_e
755: \Delta^2}{12 \pi^2 H^2} \approx \phi^2_{\rm end} +
756: \frac{g^2 N N_e \, M_p^2}{4 \pi^2} ,
757: \eeq
758: %
759: where $\phi_{\rm end}^2$ is defined by the condition that the
760: fields $h_\pm$ start to roll ($\phi^2_{\rm end} \sim \Delta/g^2$)
761: or that the slow-roll parameter $\eta$ becomes order unity
762: ($\phi^2_{\rm end} \sim g^2 N M_p^2/(8 \pi^2)$, whichever is
763: reached first. It turns out that $\Delta/g^2$ is reached first if
764: $g \gsim 10^{-3}$, and it is to this case we now specialize for
765: concreteness' sake.
766:
767: With the above choices, and taking $\phi_{\rm end}^2 \ll
768: \phi_*^2$, the amplitude of scalar fluctuations becomes
769: %
770: \beq
771: \delta_s^2 = \frac{H^2}{50 \pi^2 M_p^2 \, \epsilon_*} \sim
772: \left( \frac{16 N_e \, v^4}{75 N M_p^4} \right)\, ,
773: \eeq
774: %
775: which for 60 $e$-foldings agrees with the CMB value, $\delta_s
776: \sim 10^{-5}$ if $v/M_p \sim g \sim 10^{-3}$. The slow-roll
777: parameters at horizon exit similarly become
778: %
779: \beq
780: \epsilon_* \approx \frac{g^2 N}{32 \pi^2 N_e} \, ,
781: \qquad\qquad
782: \eta_* \approx -\; \frac{1}{2 N_e} \, .
783: \eeq
784: %
785: As is easily checked, $\phi_*$ (and so also $m_*$) is less than
786: $M_p$ provided $g^2 N N_e/(4 \pi^2) < O(1)$. This is important
787: because so long as $\phi \ll M_p$ we are justified to restrict our
788: analysis to global supersymmetry, instead of supergravity.
789:
790: In this model $\epsilon_* \ll \eta_*$, so $n_s-1 \approx 2\eta
791: \sim 10^{-2}$ (if $N_e \sim 50$). This is perfectly compatible
792: with the WMAP data; for example with the choice of cosmological
793: parameters $\Omega_b = 0.046$, $\Omega_{cdm} = 0.178$,
794: $\Omega_{\Lambda} = 1-\Omega_b-\Omega_{cdm}$, $h=0.72$, $\tau=
795: 0.17$ and $n_s=0.99$ in CMBFAST \cite{cmbfast}, we obtain $\chi^2
796: = 1430$ using the WMAP likelihood code \cite{verde}. This is as
797: good a fit to the data as the WMAP preferred parameter values
798: \cite{map-params}. Moreover since $\epsilon\ll\eta$ the model is
799: not constrained by gravitational wave production, with the WMAP
800: gravity-wave constraint giving $16\epsilon = r < 1.3$
801: \cite{map-inf}, implying the weak constraint $\epsilon < 0.08$. It
802: is interesting to note that other simple models of inflation, such
803: as $\lambda\phi^4$ chaotic inflation, are close to being ruled out
804: by the CMB data due to having too large a tilt
805: \cite{map-inf,others}.
806:
807: Having established that the model is experimentally viable, let us
808: return to the issue of decoupling. The implications of decoupling
809: are most easily seen by writing the slow-roll parameters as
810: %
811: \beq
812: (2 \epsilon_*)^{1/2} \approx \frac{g^2 N \Delta^2}{8 \pi^2 \, \rho} \, \left(
813: \frac{\varphi_* \, M_p}{m^2_*} \right) \qquad \hbox{and}
814: \qquad \eta_* \approx -\,\frac{g^2 N \Delta^2}{8 \pi^2 \,
815: \rho} \, \left(\frac{M_p^2}{m^2_*} \right) \,,
816: \eeq
817: %
818: which shows that both are suppressed by two powers of the heavy
819: mass $m_*$. What is important is that the scale against which
820: $m^2_*$ is compared is not $H^2$, but is instead $(g^2 N
821: \Delta^2/8\pi^2 \rho) M_p^2 \sim (g^4 N/8 \pi^2) M_p^2$. As we
822: dial the parameters of the model so that $m_*$ is becomes larger,
823: the slow roll parameters get closer to the scale-invariant point
824: $\epsilon_* = \eta_* = 0$. This is a consequence of decoupling,
825: since the entire inflaton potential for the model is generated by
826: virtual effects of the heavy physics. But because the benchmark
827: for observability is {\it not} $H^2/m^2_*$, the difference from
828: scale invariance can be kept observable even if $H^2/m_*^2$ is
829: much smaller than a few percent.
830:
831: In particular, we can consider two models which differ only in
832: their spectrum of particles at scale $m_*$, by comparing the
833: effects two theories which differ only in the number of heavy
834: multiplets. We take $N = 1$ for model 1 and $N=2$ for model 2, and
835: we suppose both models to undergo the same number of $e$-foldings
836: of inflation, $N_e$. With these choices these models agree on
837: their predictions for $\eta_*$. They also predict identical
838: fluctuation amplitudes provided $v_1^4 = v_2^4/2$, and so if they
839: further share the same couplings, $g_1 \sim g_2$, then we have
840: $\epsilon_{*2} = 2 \epsilon_{*1}$. Clearly the two models
841: therefore can have detectable differences in their predictions for
842: CMB observables, provided only that $\epsilon_*$ is larger than
843: $O(0.01)$.
844:
845: \subsection{Supergravity Models}\label{subsec:sugra}
846: %
847: This section makes a technical point by showing that a model
848: similar to the previous hybrid inflation model can be embedded
849: into supergravity without losing its main features. We do so
850: because we have seen that inflationary applications can (but need
851: not, for $g \ll 1$) require fields $|\phi| \gg M_p$, for which
852: supergravity corrections to the action are generally not
853: negligible.
854:
855: The ability to make the extension from global to local
856: supersymmetry is not trivial, and would have been much more
857: difficult if we had adopted an example relying purely on the
858: scalar potential's $F$-terms. In the $F$-term case the
859: supergravity corrections typically introduce new terms which are
860: of order $H^2 |\phi|^2$, and so which contribute an ${\cal O}(1)$
861: amount to the slow-roll parameter $\eta$ --- a result known as the
862: supersymmetric $\eta$-problem \cite{etaproblem}. Our main purpose
863: here is to show how this problem is evaded in the case of $D$-term
864: inflation \cite{Dterm}.
865:
866: Recall for these purposes the form of the supergravity scalar
867: sector \cite{sugra}. Given the K\"ahler function, $K(z,z^*)$, and
868: superpotential, $W(z)$, the kinetic and potential terms for a
869: collection of chiral scalars, $z^i$, are:
870: %
871: \beq
872: {\cal L} = - \sqrt{-g} \left[ \frac12 \, K_{i}^{j^*}(z,z^*) \,
873: \partial_\mu z^i \partial^\mu z^*_j + V(z,z^*) \right] \, ,
874: \eeq
875: %
876: and $V = V_F + V_D$ with
877: %
878: \beq
879: V_F = e^{K/M_p^2} \, \left[ \hat{K}^i_{j^*} (D_i W) (D^j W)^* -
880: \frac{3 \, |W|^2}{M_p^2} \right] \, ,
881: \eeq
882: %
883: and
884: %
885: \beq
886: V_D = \frac12 \, f^{ab} \, \Bigl[ K_i (T_a z)^i + \xi_a \Bigr]
887: \, \Bigl[ K_j (T_b z)^j + \xi_b \Bigr] \, .
888: \eeq
889: %
890: Here $K_i^{j^*} = \partial^2 K/\partial z^i \partial z^*_j$ and
891: $\hat{K}^i_{j^*}$ is its inverse matrix. Similarly $K_i = \partial
892: K/\partial z^i$, $T_b$ is a gauge-group generator, $\xi_a$ is a
893: Fayet-Iliopoulos term (and so is only present for $U(1)$
894: gauge-group factors) and $f^{ab}$ is the inverse matrix of Re
895: $F_{ab}(z)$, where $F_{ab}$ is the holomorphic gauge-kinetic
896: function. In these expressions the K\"ahler derivative is
897: %
898: \beq
899: D_i W = \frac{\partial W}{\partial z^i} + \frac{W}{M_p^2}
900: \frac{\partial K}{\partial z^i} = W_i + \frac{K_i \, W}{M_p^2} \,
901: .
902: \eeq
903:
904: We now specialize as before to an inflaton multiplet, $\Phi$, plus
905: two electrically charged multiplets, $H_\pm$. Write $\varphi =
906: (\Phi + \Phi^*)/M_p$ and take the no-scale ansatz \cite{noscale}
907: %
908: \beq
909: K(\varphi,\Hp^* \Hp, \Hm^* \Hm) = - 3 \, M_p^2 \, \log\varphi +
910: k(\Hp^* \Hp) + \kbr(\Hm^*\Hm) \, ,
911: \eeq
912: %
913: $F_{ab} = \delta_{ab}$ and $W = W(\Hp,\Hm)$. With these choices we
914: have $W_\Phi = \partial W/\partial \Phi = 0$ and the K\"ahler
915: derivatives are:
916: %
917: \beqa
918: D_\Phi W &=& - \, \frac{3 \, W}{M_p \, \varphi} \nonumber\\
919: D_{\Hp} W &=& W_{\Hp} + {k' \Hp^* \, W \over M_p^2} \, , \\
920: D_{\Hm} W &=& W_{\Hm} + {\kbr' \Hm^* \, W \over M_p^2} \, ,
921: \eeqa
922: %
923: where $k' = \partial k/\partial x$, for $x = \Hp^* \Hp$, and so
924: on. For instance, if $k = \Hp^* \Hp$ then $k' = 1$ {\it etc.}
925:
926: The K\"ahler metric is
927: %
928: \beq
929: K_i^{j^*} = \pmatrix{ 3/\varphi^2 & 0 & 0 \cr 0 & k' + k''
930: \Hp^*\Hp
931: & 0 \cr 0 & 0 & \kbr' + \kbr'' \Hm^* \Hm \cr} \,
932: \eeq
933: %
934: and so its inverse matrix becomes
935: %
936: \beq
937: \hat{K}^i_{j^*} =
938: \pmatrix{\varphi^2/3 & 0 & 0 \cr 0 & 1/(k' + k'' \Hp^*\Hp) & 0 \cr
939: 0 & 0 & 1/(\kbr' + \kbr'' \Hm^* \Hm )\cr } \, .
940: \eeq
941: %
942: Positivity of the kinetic energies requires $k' + k'' \Hp^*\Hp >
943: 0$ and $\kbr' + \kbr'' \Hm^* \Hm > 0$.
944:
945: With these choices the scalar potential, $V = V_F + V_D$, becomes
946: %
947: \beq
948: V_F = \frac{e^{(k + \kbr)/M_p^2}}{\varphi^3} \,
949: \left[ {|D_{\Hp} W|^2 \over k' + k'' \Hp^*\Hp}
950: + {|D_{\Hm} W|^2 \over \kbr' + \kbr'' \Hm^* \Hm} \right] \,
951: ,
952: \eeq
953: %
954: and
955: %
956: \beq
957: V_D = \frac{e^2}{2} \left[ k'|\Hp|^2 - \kbr' |\Hm|^2 + \xi \right]^2\,,
958: \eeq
959: %
960: where we take a $U(1)$ gauge group with Fayet-Iliopoulos term
961: $\xi$.
962:
963: For inflationary purposes, we may further specialize to
964: %
965: \beq
966: W = M \, (\Hp \Hm - v^2) \, ,
967: \eeq
968: %
969: and so $W_{\Hp} = M \Hm$ and $W_{\Hm} = M \Hp$. Also suppose both
970: $k'$ and $\kbr'$ are finite and nonzero as $H_\pm \to 0$, and that
971: both $k''$ and $\kbr''$ are also finite (but possibly zero) in
972: this limit. In this case we have $D_{\Hp} W \to 0$ and $D_{\Hm} W
973: \to 0$ as $H_\pm \to 0$, and so also $V_F \to 0$ in this limit.
974:
975: Any zero of both $V_F$ and $V_D$ is a global minimum for $V$,
976: which in our case is obtained by choosing $k' |\Hp|^2 + \xi =
977: \kbr' |\Hm|^2$ to ensure $V_D = 0$, and then choosing $\varphi \to
978: \infty$ to make $V_F = 0$. (This solution exists, for instance,
979: for the minimal case $k' = \kbr' = 1$.) In this limit
980: supersymmetry is broken (since $D_iW \ne 0$) but with $V = 0$
981: classically in the limit of large $\varphi$.
982:
983: There is also a trough with nonzero $V$ which is independent of
984: $\phi$, corresponding to the case $H_\pm = 0$. In this limit we
985: have $D_{H_\pm} W = 0$, and so $V_F = 0$, and so $V = V_D = e^2
986: \xi^2/2$. The mass of the $H_\pm$ scalar fields about this trough
987: are both positive and of order $M$, provided that $v \ll M_p$ and
988: $M \gg e \, \sqrt\xi$. This trough is the direct analogue of the
989: trough considered above in the globally-supersymmetric case, and
990: we may apply a similar analysis again here to the same effect.
991: Just as for the globally-supersymmetric case we have a classically
992: flat potential, whose degeneracy gets lifted by virtual loops of
993: heavy particles, and so for which the relative effect of a heavy
994: multiplet for the slow-roll parameters need not be small.
995:
996: The effective potential for this SUGRA model is similar to that of
997: the previous SUSY model, with an important qualitative difference.
998: Here the field-dependent scalar field masses for $H_\pm$ take the
999: form
1000: %
1001: \beq
1002: M^2\pm = {M^2\over \phi^3}\mp e^2\xi
1003: \eeq
1004: %
1005: while their superpartners have masses $m^2_\pm = {M^2\over
1006: \phi^3}$. In the limit where ${M^2\over \phi^3}\gg e^2\xi$, the
1007: effective potential is approximately
1008: %
1009: \beq
1010: V_{\rm eff} = \frac12{e^2}\xi^2 \left(1 - {3e^2\over
1011: 8\pi^2}\ln{\phi\over\phi_+}\right)
1012: \eeq
1013: %
1014: Therefore $\phi$ rolls to larger instead of smaller values, which
1015: is expected since the global minimum is at $\phi\to\infty$ in this
1016: model. Unlike the SUSY model, the slow roll approximation only
1017: gets better as the inflaton rolls, so the end of inflation is
1018: definitely triggered by the instability of $H_+$, at a field value
1019: given by
1020: %
1021: \beq
1022: \phi_e^3 = {M^2\over e^2\xi}
1023: \eeq
1024: %
1025: The smallest allowable initial value of $\phi$ is determined by
1026: the observational constraint on $n_s-1 \cong 2\eta\lsim 0.2$,
1027: %
1028: \beq
1029: \phi_0^2 = {3e^2\over 8\pi^2 \eta}
1030: \eeq
1031: %
1032: and the relation between $\phi$ and the number of e-foldings,
1033: $\phi^2 = \phi_0^2 + {3e^2\over 4\pi^2} N_e$, gives the
1034: constraint that
1035: %
1036: \beq
1037: \label{e5}
1038: \phi_e^2 - \phi_0^2 = {3e^2\over 4\pi^2} N_e \quad\to \quad
1039: e^5 = {M^2\over \xi}
1040: \left( {1\over 4\pi^2}\left(N_e + {3\over 2\eta}\right)\right)^{-3/2}
1041: \eeq
1042: %
1043: The COBE normalization implies
1044: %
1045: \beq
1046: \xi = 5\times 10^{-4}{\sqrt{3\eta}\over 4\pi} M_p^2
1047: \eeq
1048: %
1049: independently of the coupling $e^2$; thus $\xi$ is safely below
1050: the Planck scale. In fact, this model has the possiblity of
1051: choosing $M$ and $\phi_e$ to be subPlanckian if desired, since
1052: $M^2$ is freely adjustable. If we choose $M$ such that $\phi_e
1053: <1$, then \pref{e5} implies that $e^3 <
1054: (4\pi^2/(N_e+3/2\eta))^{3/2}$, which is a weak constraint on the
1055: coupling. Even though the global minimum is at the superPlanckian
1056: value $\phi\to\infty$, all the inflationary dynamics can be
1057: comfortably accomplished in the regime where $\phi < 1$.
1058:
1059: \section{Conclusions}\label{sec:conclusions}
1060:
1061: What we hope to have made clear is that some care must be taken
1062: when trying to use decoupling to limit the influence of heavy
1063: physics on low energy observables. We have argued that for a class
1064: of inflationary models, the inflaton potential is \emph{solely}
1065: due to the effects of integrating out the heavy particles of the
1066: model. This allows for changes in the CMB parameters of order
1067: unity, instead of the $H^2\slash M^2$ effects that might na\"ively
1068: be expected. Together with the non-adiabatic mechanism of
1069: ref.~\cite{us} this gives two ways for heavy physics to be
1070: relevant to the physics of horizon exit, using only garden-variety
1071: hybrid inflation models.
1072:
1073: Although our calculations are purely sub-Planckian, our
1074: conclusions do allow some inferences concerning trans-Planckian
1075: physics. If the trans-Planckian physics decouples, it may be able
1076: to take advantage of the same kinds of mechanisms we have found to
1077: leave some residue of itself in the CMB temperature anisotropies.
1078: If, on the other hand, the trans-Planckian physics does not
1079: decouple (as the $\alpha$-vacua proposals may not do) then its
1080: low-energy implications are likely to be everywhere, and the first
1081: problem is really to understand why low-energy predictions have
1082: been possible at all to this point, before worrying specifically
1083: about its implications for the CMB.
1084:
1085: \acknowledgments
1086:
1087: We would like to acknowledge the Aspen Center for Physics where
1088: this work was begun. We thank Robert Brandenberger, Brian Greene,
1089: Nemanja Kaloper, Anupam Mazumdar and Steve Shenker for helpful
1090: discussions. R.~H. was supported in part by DOE grant
1091: DE-FG03-91-ER40682, while C.B. and J.C. partially supported by
1092: grants from McGill University, N.S.E.R.C. (Canada) and F.C.A.R.
1093: (Qu\'ebec).
1094:
1095: \begin{thebibliography}{99}
1096:
1097: \bibitem{boomerang}http://www.physics.ucsb.edu/~boomerang/
1098:
1099: \bibitem{dasi}http://astro.uchicago.edu/dasi/
1100:
1101: \bibitem{WMAP}
1102: http://map.gsfc.nasa.gov
1103:
1104: \bibitem{Planck} http://astro.estec.esa.nl/Planck/
1105:
1106: \bibitem{tp1}
1107: %\cite{Martin:2000xs}
1108: %\bibitem{Martin:2000xs}
1109: J.~Martin and R.~H.~Brandenberger, ``The trans-Planckian problem
1110: of inflationary cosmology,'' Phys.\ Rev.\ D {\bf 63}, 123501
1111: (2001) [arXiv:hep-th/0005209];
1112: %%CITATION = HEP-TH 0005209;%%
1113: %\cite{Brandenberger:2000wr}
1114: %\bibitem{Brandenberger:2000wr}
1115: R.~H.~Brandenberger and J.~Martin, ``The robustness of inflation
1116: to changes in super-Planck-scale physics,'' Mod.\ Phys.\ Lett.\ A
1117: {\bf 16}, 999 (2001) [arXiv:astro-ph/0005432];
1118: %%CITATION = ASTRO-PH 0005432;%%
1119: %
1120: R.~H.~Brandenberger and J.~Martin, ``On the Dependence of the
1121: Spectra of Fluctuations in Inflationary Cosmology on
1122: Trans-Planckian Physics'' [arXiv:hep-th/0305161].
1123:
1124: \bibitem{tp2}
1125: %\cite{Easther:2001fi}
1126: %\bibitem{Easther:2001fi}
1127: R.~Easther, B.~R.~Greene, W.~H.~Kinney and G.~Shiu, ``Inflation as
1128: a probe of short distance physics,'' Phys.\ Rev.\ D {\bf 64},
1129: 103502 (2001) [arXiv:hep-th/0104102];
1130: %%CITATION = HEP-TH 0104102;%%
1131: %\cite{Easther:2001fz}
1132: %\bibitem{Easther:2001fz}
1133: R.~Easther, B.~R.~Greene, W.~H.~Kinney and G.~Shiu, ``Imprints of
1134: short distance physics on inflationary cosmology,''
1135: [arXiv:hep-th/0110226];
1136: %%CITATION = HEP-TH 0110226;%%
1137: %\cite{Easther:2002xe}
1138: %\bibitem{Easther:2002xe}
1139: R.~Easther, B.~R.~Greene, W.~H.~Kinney and G.~Shiu, ``A generic
1140: estimate of trans-Planckian modifications to the primordial power
1141: spectrum in inflation,'' Phys.\ Rev.\ D {\bf 66}, 023518 (2002)
1142: [arXiv:hep-th/0204129];
1143: %%CITATION = HEP-TH 0204129;%%
1144:
1145: \bibitem{tp3}
1146: %\cite{Chu:2000ww}
1147: %\bibitem{Chu:2000ww}
1148: C.~S.~Chu, B.~R.~Greene and G.~Shiu, ``Remarks on inflation and
1149: noncommutative geometry,'' Mod.\ Phys.\ Lett.\ A {\bf 16}, 2231
1150: (2001) [arXiv:hep-th/0011241];
1151: %%CITATION = HEP-TH 0011241;%%
1152: %\cite{Martin:2000bv}
1153: %\bibitem{Martin:2000bv}
1154: J.~Martin and R.~H.~Brandenberger, ``A cosmological window on
1155: trans-Planckian physics,'' [arXiv:astro-ph/0012031];
1156: %%CITATION = ASTRO-PH 0012031;%%
1157: %\cite{Tanaka:2000jw}
1158: %\bibitem{Tanaka:2000jw}
1159: T.~Tanaka, ``A comment on trans-Planckian physics in inflationary
1160: universe,'' [arXiv:astro-ph/0012431];
1161: %%CITATION = ASTRO-PH 0012431;%%
1162: %\cite{Niemeyer:2001qe}
1163: %\bibitem{Niemeyer:2001qe}
1164: J.~C.~Niemeyer and R.~Parentani, ``Trans-Planckian dispersion and
1165: scale-invariance of inflationary perturbations,'' Phys.\ Rev.\ D
1166: {\bf 64}, 101301 (2001) [arXiv:astro-ph/0101451];
1167: %%CITATION = ASTRO-PH 0101451;%%
1168: %\cite{Kempf:2001fa}
1169: %\bibitem{Kempf:2001fa}
1170: A.~Kempf and J.~C.~Niemeyer, ``Perturbation spectrum in inflation
1171: with cutoff,'' Phys.\ Rev.\ D {\bf 64}, 103501 (2001)
1172: [arXiv:astro-ph/0103225];
1173: %%CITATION = ASTRO-PH 0103225;%%
1174: %\cite{Starobinsky:2001kn}
1175: %\bibitem{Starobinsky:2001kn}
1176: A.~A.~Starobinsky, ``Robustness of the inflationary perturbation
1177: spectrum to trans-Planckian physics,'' Pisma Zh.\ Eksp.\ Teor.\
1178: Fiz.\ {\bf 73}, 415 (2001) [JETP Lett.\ {\bf 73}, 371 (2001)]
1179: [arXiv:astro-ph/0104043];
1180: %%CITATION = ASTRO-PH 0104043;%%
1181: %\cite{Bastero-Gil:2001js}
1182: %\bibitem{Bastero-Gil:2001js}
1183: M.~Bastero-Gil, ``What can we learn by probing trans-Planckian
1184: physics,'' [arXiv:hep-ph/0106133];
1185: %%CITATION = HEP-PH 0106133;%%
1186: %\cite{Lemoine:2001ar}
1187: %\bibitem{Lemoine:2001ar}
1188: M.~Lemoine, M.~Lubo, J.~Martin and J.~P.~Uzan, ``The stress-energy
1189: tensor for trans-Planckian cosmology,'' Phys.\ Rev.\ D {\bf 65},
1190: 023510 (2002) [arXiv:hep-th/0109128];
1191: %%CITATION = HEP-TH 0109128;%%
1192: %\cite{Brandenberger:2002ty}
1193: %\bibitem{Brandenberger:2002ty}
1194: R.~H.~Brandenberger, S.~E.~Joras and J.~Martin, ``Trans-Planckian
1195: physics and the spectrum of fluctuations in a bouncing universe,''
1196: [arXiv:hep-th/0112122];
1197: %%CITATION = HEP-TH 0112122;%%
1198: %\cite{Martin:2002kt}
1199: %\bibitem{Martin:2002kt}
1200: J.~Martin and R.~H.~Brandenberger, ``The Corley-Jacobson
1201: dispersion relation and trans-Planckian inflation,'' Phys.\ Rev.\
1202: D {\bf 65}, 103514 (2002) [arXiv:hep-th/0201189];
1203: %%CITATION = HEP-TH 0201189;%%
1204: %\cite{Danielsson:2002kx}
1205: %\bibitem{Danielsson:2002kx}
1206: U.~H.~Danielsson, ``A note on inflation and trans-Planckian
1207: physics,'' Phys.\ Rev.\ D {\bf 66}, 023511 (2002)
1208: [arXiv:hep-th/0203198];
1209: %%CITATION = HEP-TH 0203198;%%
1210: %\cite{Hassan:2002qk}
1211: %\bibitem{Hassan:2002qk}
1212: S.~F.~Hassan and M.~S.~Sloth, ``Trans-Planckian effects in
1213: inflationary cosmology and the modified uncertainty principle,''
1214: [arXiv:hep-th/0204110];
1215: %%CITATION = HEP-TH 0204110;%%
1216: %\cite{Danielsson:2002qh}
1217: %\bibitem{Danielsson:2002qh}
1218: U.~H.~Danielsson, ``Inflation, holography and the choice of vacuum
1219: in de Sitter space,'' JHEP {\bf 0207}, 040 (2002)
1220: [arXiv:hep-th/0205227];
1221: %%CITATION = HEP-TH 0205227;%%
1222: A.~A.~Starobinsky and I.~I.~Tkachev,
1223: %``Trans-Planckian particle creation in cosmology and ultra-high energy cosmic rays,''
1224: JETP Lett.\ {\bf 76}, 235 (2002) [Pisma Zh.\ Eksp.\ Teor.\ Fiz.\
1225: {\bf 76}, 291 (2002)] [arXiv:astro-ph/0207572];
1226: %%CITATION = ASTRO-PH 0207572;%%
1227: %\cite{Goldstein:2002fc}
1228: %\bibitem{Goldstein:2002fc}
1229: K.~Goldstein and D.~A.~Lowe, ``Initial state effects on the cosmic
1230: microwave background and trans-planckian physics,''
1231: [arXiv:hep-th/0208167];
1232: %%CITATION = HEP-TH 0208167;%%
1233: %\cite{Danielsson:2002mb}
1234: %\bibitem{Danielsson:2002mb}
1235: U.~H.~Danielsson, ``On the consistency of de Sitter vacua,''
1236: [arXiv:hep-th/0210058];
1237: %%CITATION = HEP-TH 0210058;%%
1238: %\cite{Shiu:2002kg}
1239: G.~Shiu and I.~Wasserman, ``On the signature of short distance
1240: scale in the cosmic microwave background,'' Phys.\ Lett.\ B {\bf
1241: 536}, 1 (2002) [arXiv:hep-th/0203113];
1242: %%CITATION = HEP-TH 0203113;%%
1243:
1244: \bibitem{shenker}
1245: N.~Kaloper, M.~Kleban, A.~Lawrence,
1246: S.~Shenker, ``Signatures of short distance physics in the cosmic
1247: microwave background'', [arXiv:0201158];
1248: %
1249: N.~Kaloper, M.~Kleban,
1250: A.~Lawrence, S.~Shenker, and L.~Susskind,
1251: ``Initial conditions for inflation'', [arXiv:hep-th/0209231].
1252:
1253: \bibitem{dSinconsistent}
1254: T. Banks and L. Mannelli, ``De Sitter vacua, renormalization and
1255: locality'', Phys.\ Rev.\ {\bf D67} {2003} {065009}, [arXiv:hep-th/0209113];
1256: %
1257: M. Einhorn and F. Larsen, ``Interacting quantum field theory in de
1258: Sitter vacua'', Phys. \ Rev. \ {\bf D67} {2003} {024001}, [arXiv:hep-th/0209159];
1259: %
1260: M. Einhorn and F. Larsen, ``Squeezed States in the de Sitter Vacuum'', [arXiv: hep-th/0305056];
1261: %
1262: K.~Goldstein and D.~A.~Lowe, ``A note on alpha-vacua and
1263: interacting field theory in de Sitter space'' ,[arXiv:
1264: hep-th/0302050].
1265: %
1266: Hael Collins, R.~Holman and Matthew~R.~Martin, ``The Fate of the
1267: $\alpha$-Vacuum'', [arXiv:hep-th/0306028].
1268:
1269: \bibitem{WMAPinflation}
1270: H.V. Peiris, {\it et. al.}, [arXiv:astro-ph/0302225];
1271: %
1272: V. Barger, H.-S. Lee and D. Marfatia, [arXiv:hep-ph/0302150];
1273: %
1274: B. Kyae and Q. Shafi, [arXiv:astro-ph/0302504];
1275: %
1276: J.R. Ellis, M. Raidal and T. Yanagida, [arXiv:hep-ph/0303242].
1277:
1278: \bibitem{us}
1279: C.P. Burgess, J.M. Cline, F. Lemieux and R. Holman, JHEP 0302
1280: (2003) 048 (24 pages) [ArXiv: hep-th/0210233].
1281: %%CITATION = HEP-TH 0210233;%%
1282:
1283: \bibitem{RobertRMP}
1284: R.H.~Brandenberger, Rev.\ Mod.\ Phys.\ {\bf 57} (1985) 1.
1285:
1286: \bibitem{hybrid}
1287: A.~Linde, Phys.\ Rev. {\bf D49} (1994) 748
1288: [arXiv:astro-ph/9307002];
1289: %
1290: David H.~ Lyth and Ewan Stewart,`` More varieties of hybrid
1291: inflation'', Phys.\ Rev.\ {\bf D54} (1996) {7186},
1292: [arXiv:hep-ph/9606412];
1293:
1294: \bibitem{liddlelyth}
1295: Andrew R.~ Liddle and David H.~ Lyth
1296: ``Cosmological Inflation and Large Scale Structure'', Cambridge
1297: University Press (2000).
1298:
1299: \bibitem{WMAPnrun}
1300: D.N. Spergel {\it et.al.}, [arXiv:astro-ph/0302209].
1301:
1302: \bibitem{BI}
1303: G.R. Dvali and S.H.Henry Tye, Phys.\ Lett.\ {\bf B450} (1999)
1304: 72-82 [arXiv:hep-ph/9812483];
1305: %
1306: A. Lukas, B.A. Ovrut and D. Waldram, Phys.\ Rev.\ {\bf D61} (2000)
1307: 023506 [arXiv:hep-th/9902071];
1308: %
1309: J.M. Cline, Phys.\ Rev.\ {\bf D61} (2000) 023513
1310: [arXiv:hep-ph/9904495];
1311: %
1312: E.E. Flanagan, S.H.Henry Tye and I. Wasserman, Phys.\ Rev.\ {\bf
1313: D62} (2000) 024011 [arXiv:hep-ph/9909373];
1314: %
1315: A. Kehagias and E. Kiritsis, JHEP (1999) 9911:022
1316: [arXiv:hep-th/9910174];
1317: %
1318: J. Khoury, B.A. Ovrut, P.J. Steinhardt and N. Turok, Phys.\ Rev.\
1319: {\bf D64} (2001) 123522 [arXiv:hep-th/0103239];
1320: %
1321: C.P. Burgess, M. Majumdar, D. Nolte, F. Quevedo and G. Rajesh,
1322: JHEP 0107 (2001) 047 [arXiv:hep-th/0105204];
1323: %%CITATION = HEP-TH 0105204;%%
1324: %
1325: C. Herdeiro, S. Hirano and R. Kallosh, JHEP 0112 (2001) 027
1326: [arXiv:hep-th/0110271];
1327: %
1328: C.P. Burgess, P. Martineau, F. Quevedo, G. Rajesh and R.J. Zhang,
1329: JHEP 0203 (2002) 052 [arXiv:hep-th/0111025];
1330: %%CITATION = HEP-TH 0111025;%%
1331: %
1332: J. Garcia-Bellido, R. Rabadan and F. Zamora, JHEP 0201 (2002) 036
1333: [arXiv:hep-th/0112147];
1334: %
1335: R. Blumenhagen, B. Kors, D. Lust and T. Ott, Nucl.\ Phys.\ {\bf
1336: B641} (2002) 235-255 [arXiv:hep-th/0202124];
1337: %
1338: K. Dasgupta, C. Herdeiro, S. Hirano and R. Kallosh, Phys.\ Rev.\
1339: {\bf D65} (2002) 126002 [arXiv:hep-th/0203019];
1340: %
1341: J.-Y. Kim, Phys.\ Lett.\ {\bf B548} (2002) 1-8
1342: [arXiv:hep-th/0203084];
1343: %
1344: N. Jones, H. Stoica and S.H.Henry Tye, JHEP 0207 (2002) 051
1345: [arXiv:hep-th/0203163].
1346:
1347:
1348: \bibitem{lythriotto} David H.~ Lyth and Antonio Riotto, ``Particle Physics
1349: Models of Inflation and the Cosmological Density Perturbation'',
1350: \prep{314}{1}{1999}.
1351:
1352: \bibitem{Fterm}
1353: M. Bastero-Gil and S.F. King, Nucl.\ Phys.\ {\bf B549} (1999)
1354: 391-406 [arXiv:hep-ph/9806477];
1355: %
1356: J.A. Casas, G.B. Gelmini and A. Riotto, Phys.\ Lett.\ {\bf B459}
1357: (1999) 91-96 [arXiv:hep-ph/9903492];
1358: %
1359: S.M. Harun-or-Rashid, T. Kobayashi and H. Shimabukuro, Phys.\
1360: Lett.\ {\bf B466} (1999) 95-99 [arXiv:hep-ph/9908266];
1361: %
1362: J. McDonald, JHEP 0212 (2002) 029 [arXiv:hep-ph/0201016].
1363:
1364: \bibitem{Dterm}
1365: P. Binetruy and G.R. Dvali, Phys.\ Lett.\ {\bf B388} (1996)
1366: 241-246, [arXiv:hep-ph/9606342];
1367: %
1368: E. Halyo, Phys.\ Lett.\ {\bf B387} (1996) 43-47
1369: [arXiv:hep-ph/9606423];
1370: %
1371: T. Matsuda, Phys.\ Lett.\ {\bf B423} (1998) 35-39
1372: [arXiv:hep-ph/9705448];
1373: %
1374: D.H. Lyth and A. Riotto, Phys.\ Lett.\ {\bf B412} (1997) 28-34
1375: [arXiv:hep-ph/9707273];
1376: %
1377: C.F. Kolda and J. March-Russell, Phys.\ Rev.\ {\bf D60} (1999)
1378: 023504 [arXiv:hep-ph/9802358];
1379: %
1380: J.R. Espinosa, A. Riotto and G.G. Ross, Nucl.\ Phys.\ {\bf B531}
1381: (1998) 461-477 [arXiv:hep-ph/9804214];
1382: %
1383: S.F. King and A. Riotto, Phys.\ Lett.\ {\bf B442} (1998) 68-73
1384: [arXiv:hep-ph/9806281];
1385: %
1386: J. Lesgourgues, Phys.\ Lett.\ {\bf B452} (1999) 15-22
1387: [arXiv:hep-ph/9811255];
1388: %
1389: E. Halyo, Phys.\ Lett.\ {\bf B454} (1999) 223-227
1390: [arXiv:hep-ph/9901302]; Phys.\ Lett.\ {\bf B461} (1999) 109-113
1391: [arXiv:hep-ph/9905244];
1392:
1393: \bibitem{gShybrid}
1394: C. Panagiotakopoulos, Phys.\ Rev.\ {\bf D55} (1997) 7335-7339
1395: [arXiv:hep-ph/9702433]; Phys.\ Lett.\ {\bf B402} (1997) 257-262
1396: [arXiv:hep-ph/9703443];
1397: %
1398: A.D. Linde and A. Riotto, Phys.\ Rev.\ {\bf D56} (1997) 1841--1844
1399: [arXiv:hep-ph/9703209];
1400: %
1401: L. Covi, G. Mangano, A. Masiero and G. Miele, Phys.\ Lett.\ {\bf
1402: B424} (1998) 253-258 [arXiv:hep-ph/9707405];
1403: %
1404: M.~Bastero-Gil and S.~F.~ King, Phys.\ Lett.\ {\bf B423} (1998)
1405: {27} [arXiv:astro-ph/9709502];
1406: %
1407: G.R. Dvali, G. Lazarides and Q. Shafi, Phys.\ Lett.\ {\bf B424}
1408: (1998) 259-264 [arXiv:hep-ph/9710314];
1409: %
1410: T. Watari and T. Yanagida, Phys.\ Lett.\ {\bf B499} (2001) 297-304
1411: [arXiv:hep-ph/0011389];
1412:
1413: \bibitem{cmbfast}U.~Seljak and M.~ Zaldarriaga, ``A Line of Sight Approach to Cosmic
1414: Microwave Background Anisotropies'', \apj{469}{1}{1996}; see also the site at
1415: http://www.sns.ias.edu/~matiasz/CMBFAST/cmbfast.html.
1416: %%CITATION = ASTRO-PH 9911219;%%
1417:
1418: \bibitem{verde}
1419: L.~Verde {\it et al.}, ``First Year Wilkinson Microwave Anisotropy
1420: Probe (WMAP) Observations: Parameter Estimation Methodology,''
1421: arXiv:astro-ph/0302218.
1422: %%CITATION = ASTRO-PH 0302218;%%
1423:
1424: \bibitem{map-params}
1425: D.~N.~Spergel {\it et al.}, ``First Year Wilkinson Microwave
1426: Anisotropy Probe (WMAP) Observations: Determination of
1427: Cosmological Parameters,'' arXiv:astro-ph/0302209.
1428: %%CITATION = ASTRO-PH 0302209;%%
1429:
1430: \bibitem{map-inf}
1431: H.~V.~Peiris {\it et al.}, ``First year Wilkinson Microwave
1432: Anisotropy Probe (WMAP) observations: Implications for
1433: inflation,'' arXiv:astro-ph/0302225.
1434: %%CITATION = ASTRO-PH 0302225;%%
1435:
1436: \bibitem{others}
1437: V.~Barger, H.~S.~Lee and D.~Marfatia,
1438: ``WMAP and inflation,''
1439: arXiv:hep-ph/0302150;\\
1440: %%CITATION = HEP-PH 0302150;%%
1441: S.~Dodelson and L.~Hui,
1442: ``A horizon ratio bound for inflationary fluctuations,''
1443: arXiv:astro-ph/0305113;\\
1444: W.~H.~Kinney, E.~W.~Kolb, A.~Melchiorri and A.~Riotto,
1445: ``WMAPping inflationary physics,''
1446: arXiv:hep-ph/0305130;\\
1447: %%CITATION = HEP-PH 0305130;%%
1448: %%CITATION = ASTRO-PH 0305113;%%
1449: A.~R.~Liddle and S.~M.~Leach, ``How long before the end of
1450: inflation were observable perturbations produced?,''
1451: arXiv:astro-ph/0305263.
1452: %%CITATION = ASTRO-PH 0305263;%%
1453:
1454: \bibitem{etaproblem}
1455: E.J. Copeland, A.R. Liddle, D.H. Lyth, E.D. Stewart and D. Wands,
1456: Phys.\ Rev.\ {\bf D49} (1994) 6410 [arXiv:astro-ph/9401011];
1457: %
1458: For a recent discussion see, N. Arkani-Hamed, H.-C. Cheng, P.
1459: Creminelli and L. Randall, [arXiv:hep-th/0302034].
1460:
1461: \bibitem{sugra}
1462: E. Cremmer, {\it et. al.}, Nucl.\ Phys.\ {\bf B147} (1979) 105;
1463: %
1464: E. Witten and J. Bagger, Phys.\ Lett.\ {\bf B115} (1982) 202.
1465:
1466: \bibitem{noscale}
1467: J.R. Ellis, K. Enqvist and D.V. Nanopoulos, Phys.\ Lett.\ {\bf
1468: B147} (1984) 99;
1469: %
1470: J.R. Ellis, C. Kounnas and D.V. Nanopoulos, Nucl.\ Phys.\ {\bf B
1471: 241} (1984) 406.
1472:
1473:
1474: \end{thebibliography}
1475:
1476: \end{document}
1477: