1: %===========================================================================
2: % IR Divergences in dS Vers: 1.0 |
3: %===========================================================================
4:
5:
6: %===========================================================================
7: % PREAMBLE |
8: %===========================================================================
9:
10:
11: \documentclass[12pt,letterpaper]{JHEP3}
12:
13:
14: %------------------------------ PACKAGES -----------------------------------
15:
16:
17:
18: \usepackage{amssymb}
19: \usepackage{amsmath,amsfonts}
20: \usepackage{graphicx}
21: %\usepackage{axodraw}
22: %\usepackage{epsfig}
23:
24:
25: \def\KeyWord#1{$\backslash$\IfColor{$\!\!$\textRed{#1}\textBlack}{#1}$\!\!$}
26:
27:
28:
29: %------------------------------ DEFINITIONS --------------------------------
30:
31: \providecommand{\slashD}{D\mspace{-12mu}\slash\mspace{3mu}}
32: \providecommand{\slashdel}{\partial\mspace{-9mu}\slash}
33: \providecommand{\sgn}{\mathrm{sgn}\,}
34: \def\Dsl{D \!\!\!\! /}
35:
36:
37:
38:
39: \def\fnote#1#2{\begingroup\def\thefootnote{#1}\footnote{#2}\addtocounter
40: {footnote}{-1}\endgroup}
41:
42:
43:
44: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
45: %%%%%%%%%%%%%%%%%%%%%%%%%%%
46: % Input macros
47: %\input psfig
48: % Equation lines without numbers
49: \def\beq{\begin{eqnarray}}
50: \def\eeq{\end{eqnarray}}
51: \def\bea{\begin{eqnarray*}}
52: \def\eea{\end{eqnarray*}}
53:
54: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
55:
56: % References to main physics journals
57: \def\NPB#1#2#3{Nucl. Phys. {\bf B#1}, #3 (#2)}
58: \def\PLB#1#2#3{Phys. Lett. {\bf B#1}, #3 (#2)}
59: \def\PLBold#1#2#3{Phys. Lett. {\bf #1B}, #3 (#2)}
60: \def\PRP#1#2#3{Phys. Rep. {\bf #1}, #3 (#2)}
61: \def\PRD#1#2#3{Phys. Rev. {\bf D#1}, #3 (#2)}
62: \def\PRold#1#2#3{Phys. Rev. {\bf #1} (#2) #3}
63: \def\PRL#1#2#3{Phys. Rev. Lett. {\bf #1}, #3 (#2)}
64: \def\PREP#1#2#3{Phys. Rep. {\bf #1} #3, (#2)}
65: \def\ZPC#1#2#3{Z. Phys. C {\bf #1}, #3 (#2)}
66:
67: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
68:
69: % \gsim and \lsim provide >= and <= signs.
70: \def\centeron#1#2{{\setbox0=\hbox{#1}\setbox1=\hbox{#2}\ifdim
71: \wd1>\wd0\kern.5\wd1\kern-.5\wd0\fi
72: \copy0\kern-.5\wd0\kern-.5\wd1\copy1\ifdim\wd0>\wd1
73: \kern.5\wd0\kern-.5\wd1\fi}}
74: \def\ltap{\;\centeron{\raise.35ex\hbox{$<$}}{\lower.65ex\hbox{$\sim$}}\;}
75: \def\gtap{\;\centeron{\raise.35ex\hbox{$>$}}{\lower.65ex\hbox{$\sim$}}\;}
76: \def\gsim{\mathrel{\gtap}}
77: \def\lsim{\mathrel{\ltap}}
78:
79: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
80:
81: \def\doublespaced{\baselineskip=\normalbaselineskip\multiply
82: \baselineskip by 200\divide\baselineskip by 100}
83: \def\singleandhalfspaced{\baselineskip=\normalbaselineskip\multiply
84: \baselineskip by 150\divide\baselineskip by 100}
85: \def\singleandabitspaced{\baselineskip=\normalbaselineskip\multiply
86: \baselineskip by 120\divide\baselineskip by 100}
87: \def\singleandthirdspaced{\baselineskip=\normalbaselineskip\multiply
88: \baselineskip by 130\divide\baselineskip by 100}
89: \def\singlespaced{\baselineskip=\normalbaselineskip}
90:
91: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
92:
93: % Various particle physics / supersymmetry commands.
94: \newcommand{\newc}{\newcommand}
95: \newc{\qbar}{{\overline q}}
96: \newc{\Kahler}{K\"ahler }
97: \newc{\deltaGS}{\delta_{\rm GS}}
98: %%%%%%%%%%%%%%%%%%%%%%%%%%
99:
100: %------------------------------ TITLE & ABSTRACT ---------------------------
101:
102: \title{Infrared Divergences in dS/CFT}
103:
104: \author{T. Banks \\
105: Department of Physics and Astronomy - NHETC\\
106: Rutgers University, Piscataway, NJ 08540\\
107: and\\
108: Department of Physics, SCIPP\\
109: University of California, Santa Cruz, CA 95064\\
110: E-mail: \email{banks@scipp.ucsc.edu}}
111:
112: \author{L. Mannelli \\
113: Department of Physics, SCIPP\\
114: University of California, Santa Cruz, CA 95064\\
115: and\\
116: Kavli Institute for Theoretical Physics\\
117: University of California\\
118: Santa Barbara, CA 93106\\
119: E-mail: \email{lorenzo@scipp.ucsc.edu}}
120:
121: \author{W. Fischler, \\
122: Department of Physics\\
123: University of Texas, Austin, TX 78712\\
124: E-mail: \email{fischler@utexas.edu}}
125:
126:
127:
128: \abstract{ dS/CFT gives a perturbatively gauge invariant definition of
129: particle masses in de Sitter (dS) space. We show, in a toy model in which
130: the graviton is replaced with a minimally coupled massless scalar field,
131: that loop corrections to these masses are infrared (IR) divergent. We
132: argue that this implies anomalous dependence of masses on the cosmological
133: constant, in a true theory of quantum gravity. This is in accord with the
134: hypothesis of Cosmological SUSY Breaking (CSB).}
135:
136: \keywords{infrared divergences, de Sitter space}
137:
138: \preprint{\hepth{} \\SCIPP 05/31\\ UTTG-02-05}
139:
140:
141: \begin{document}
142:
143:
144: %===========================================================================
145: % BULK PAPER |
146: %===========================================================================
147:
148: % Table of contents automatic !!!
149:
150: \section{\textbf{Introduction}}
151:
152: The hypothesis of Cosmological Supersymmetry Breaking (CSB) is based on the
153: idea \cite{Banks:2000fe} \cite{Banks:2001zj} \cite{Fischler:2000} that
154: quantum theories of stable, asymptotically de Sitter (AsdS) space-times
155: exist and have a finite number of physical states. The (positive)
156: cosmological constant, $\Lambda$,is an input parameter, which controls the
157: number of states. The limit of vanishing $\Lambda$ is a super-Poincare
158: invariant theory, but SUSY is broken for finite $\Lambda$: the operator
159: which converges to the Poincare Hamiltonian $P_0$, does not commute with the
160: SUSY charges.
161:
162: Classical SUGRA supports such a picture, but suggests a relation between the
163: gravitino mass and the c.c.: $m_{3/2} \sim \Lambda^{1/2} / M_P$. CSB is the
164: proposal that the exponent $1/2$ in this relation is replaced by $1/4$ in
165: the quantum theory. Given the interpretation of $\Lambda$ as a parameter
166: controlling the number of states, this is a critical exponent, and it is
167: plausible that it has fluctuation corrections. Indeed, low energy effective
168: field theory cannot calculate the real relation between the gravitino mass
169: and the c.c., since the c.c. is a relevant parameter and one must introduce
170: a counterterm for it. The exponent above is just the ``natural" relation of
171: classical SUGRA, without fine tuning of the constant in the superpotential.
172: If we accept such fine tuning, we can get any relation we want between $%
173: m_{3/2}$ and $\Lambda$ in effective field theory.
174:
175: However, the necessity of canceling an infinite c.c. appears to be a short
176: distance problem in effective field theory, and as such, does not seem to
177: depend on the value of the c.c. . As a consequence, there has been
178: considerable skepticism about CSB.
179:
180: In \cite{Banks:2002wj}, one of the authors presented an argument for the
181: exponent $1/4$, based on crude approximations to the dynamics of the
182: cosmological horizon in the static observer gauge for dS space. It was clear
183: that from the static observer's point of view, the enhanced exponent is an
184: IR effect. However, since the argument relied on conjectures about the
185: horizon dynamics, it has not convinced anyone. Skeptical observers want to
186: understand where effective field theory reasoning breaks down. The work of
187: \cite{Banks:2002wr} provided an important clue. In the static gauge most of
188: the states in a quantum theory of dS space live on the horizon of the static
189: observer. Local field theory can describe only a negligible fraction of the
190: entropy. On the contrary, it was argued in \cite{Banks:2002wr} that in
191: global coordinates, the entire Hilbert space may be well described by field
192: theory. The contradiction between a finite number of states and the field
193: theoretic description can be viewed as an IR cutoff, which restricts the
194: global time coordinate to an interval of order $|t| \leq {\frac{R}{6}}
195: \mathrm{ln} \left(R M_P\right)$ around the time symmetric point. The field
196: theory also has a UV cutoff at a scale $M_c \sim \left({\frac{M_P }{R}}%
197: \right)^{1/2}$. This description is inappropriate for states containing
198: black holes whose size scales like $R$, but there is a basis of field
199: theoretic states in global coordinates, which may span the Hilbert space.
200:
201: A simple way to restate this conclusion is to invoke the fact that the
202: global description of dS space in field theory does not seem to break down
203: until we contemplate introducing black holes on early time initial data
204: slices, whose entropy exceeds that of the space-time. The combined UV and IR
205: cutoffs prevent us from introducing such objects, and describes a cutoff
206: field theory with a finite number of states. The field theory description of
207: many of these states breaks down near the time-symmetric point, but near the
208: upper and lower limits of $t$, it is a good approximation to their
209: properties.
210:
211: We have thus set up a framework in which IR divergences in a field theoretic
212: treatment of dS space can be thought of as introducing non-classical
213: dependence on the c.c. . It has often been argued that perturbative quantum
214: gravity expanded around dS space is fraught with IR divergences. These
215: claims have been less than convincing, because no-one had identified gauge
216: invariant observables with which to check the physical meaning of the
217: logarithmically growing graviton propagator. This problem is solved by
218: dS/CFT \cite{Witten:2001kn}\cite{Strominger:2001pn}\cite{Maldacena:2002vr}.
219: In particular, the mass of a field in dS space is given a gauge invariant
220: meaning: it is related to the dimension of a conformal field on the boundary.
221:
222: The plan of this paper is as follows: in the next section we review dS/CFT,
223: in the Wheeler-DeWitt formalism proposed by Maldacena. This allows
224: perturbative calculations to be performed in a straightforward manner,
225: apparently troubled only by conventional UV divergences. In section 3 we
226: perform one loop calculations of boundary dimensions in a variety of
227: non-gravitational field theories. We find that when the theory contains a
228: massless, minimally coupled scalar field with soft couplings, the dimensions
229: are infected with IR logarithms. In the conclusions, we discuss the
230: difficulties attendant on an extension of these calculations to perturbative
231: quantum gravity.
232:
233: \section{Review of dS/CFT}
234:
235: In his talk at Strings 2001 in Mumbai \cite{Witten:2001kn}, Witten proposed
236: a sort of scattering theory for de Sitter space. The fundamental object was
237: the path integral with fixed boundary conditions on $\mathcal{I}_{\pm}$. It
238: was implicitly assumed that, as in asymptotically flat and Anti-deSitter
239: spaces, a field theoretic approximation became exact near the boundaries of
240: space-time. This assumption is open to criticism. It is likely that generic
241: boundary conditions on fields on $\mathcal{I}_-$ will lead to Big Crunch
242: space-times, rather than space-times which are future asymptotically dS.
243: However, this criticism does not apply to perturbation theory, where the
244: boundary conditions are infinitesimal perturbations of those corresponding
245: to the dS vacuum. Witten's prescription provides a perturbative definition
246: of amplitudes in dS quantum gravity, which are invariant under
247: diffeomorphisms that approach the identity near $\mathcal{I}_{\pm}$.
248:
249: Somewhat later, Strominger proposed \cite{Strominger:2001pn} that suitably
250: defined boundary amplitudes should be the correlation functions of a
251: Euclidean conformal field theory (CFT). An apparent difference with Witten's
252: proposal is the role of conformally covariant, rather than invariant
253: amplitudes in dS/CFT. However, Maldacena \cite{maldapriv} has emphasized
254: that the operator dimensions, OPE coefficients and the like, of dS/CFT, are
255: gauge invariant observables in the sense of Witten.
256:
257: The boundary correlation functions defined by Strominger should certainly be
258: conformally invariant, but it is not clear that they should obey the axioms
259: of field theory. Analogous arguments would lead us to believe that the
260: holographic dual of linear dilaton backgrounds \cite{Aharony:1999ti} was a
261: Lorentz invariant field theory. The calculations of Peet and Polchinski \cite%
262: {Peet:1998wn} show that it is not. In the dS/CFT case, the form of the two
263: point function follows from conformal invariance alone, and does not give us
264: enough of a clue to the nature of the holographic dual. As believers in the
265: proposition that quantum dS space has only a finite number of states, the
266: present authors are inclined to disbelieve that a CFT will be the exact
267: description of the quantum theory.
268:
269: For our present purposes, all of these issues of principle are somewhat
270: beside the point. We want a definition of correlation functions on $\mathcal{%
271: I}_{\pm}$ which is perturbatively well defined and gauge invariant.
272: Furthermore, we will be interested only in two point functions, and will not
273: have to address the question of whether higher order correlators obey the
274: axioms of CFT. We have found that the dS/CFT prescription advocated by
275: Maldacena \cite{Maldacena:2002vr} is the most appropriate for our purposes.
276: Maldacena observes that the Euclidean path integral on a space with the
277: topology of a hemisphere defines a ``wave function of the universe" which is
278: a functional of fields on the boundary of the hemisphere. In leading
279: semiclassical approximation, the geometry is the section of the round sphere
280: metric
281: \begin{equation*}
282: {ds^2 = d\theta^2 + sin^2 (\theta ) d\Omega^2 }
283: \end{equation*}
284: with $0 \leq \theta \leq \theta_0$. Maldacena defines boundary correlators
285: as the expansion coefficients of the logarithm of the wave function of the
286: universe for fixed $\theta_0$. The analytic extrapolation $\theta_0
287: \rightarrow {\frac{\pi}{2}} + i t$, $ t\rightarrow\infty$ defines
288: correlation functions on $\mathcal{I}_+$. If the limiting correlation
289: functions exist, they should be covariant under the conformal group of the
290: sphere. In particular, if we work in planar coordinates on the upper
291: triangle of the dS Penrose diagram
292: \begin{equation*}
293: ds^{2}=\frac{1}{\eta ^{2}}\left( -d\eta ^{2}+d\mathbf{x}^{2}\right)
294: \end{equation*}
295: ($\mathcal{I}_+$is at $\eta = 0$) then the boundary two point function
296: should have the form $|\mathbf{x}|^{-\Delta} $. For a free scalar field of
297: mass $m^2$ this is indeed true, and the relation between mass and dimension
298: is given by
299: \begin{equation*}
300: \Delta _{\pm }=a=\frac{1}{2}\left( d-1\pm \sqrt{(d-1)^{2}-4m^{2}R^{2}}\right)
301: \end{equation*}
302: This is an analytic continuation (in the c.c.) of analogous formulas in
303: AdS/CFT. Indeed, Maldacena's proposal for the correlation functions is the
304: direct analog of the calculation of Euclidean correlation functions in
305: AdS/CFT.
306:
307: The purpose of the present paper is to compute one loop corrections to $%
308: \Delta_{\pm}$ in simple field theory models. We will see that when the
309: theory has a massless, minimally coupled scalar with soft couplings, these
310: corrections are IR divergent.
311:
312: \section{Review of QFT in dS space}
313:
314: In this section we will introduce the principal formulae of QFT in\textit{d-}%
315: dimensional de Sitter (dS$^{d}$) space, and fix our notation .
316:
317: For a more complete discussion we refer to the excellent review paper \cite%
318: {Spradlin:2001pw}.
319:
320: \subsection{Coordinate Systems}
321:
322: \textit{d-}dimensional de Sitter dS$^{d}$ can be realized as the manifold,
323: embedded in $d+1$ dimensional Minkowski $\mathcal{M}^{d,1}$ space, defined
324: by the equation%
325: \begin{equation}
326: -X_{0}^{2}+X_{1}^{2}+\cdots X_{d}^{2}=R^{2} \label{Eq for
327: Hyperboloid}
328: \end{equation}%
329: where $R$ is the de Sitter radius.
330:
331: The de Sitter metric is the standard metric induced by immersion in $%
332: \mathcal{M}^{d,1}$ with the usual flat metric. The isometry group of dS$^{d}$
333: is $O(d,1)$ in fact this leave invariant both the hyperboloid defined by the
334: equation (\ref{Eq for Hyperboloid}) and the flat metric of $\mathcal{M}%
335: ^{d,1} $.
336:
337: For the most part, we will use planar coordinates%
338: \begin{eqnarray}
339: X^{0} &=&\sinh t-\frac{1}{2}x_{i}x_{i}e^{-t} \notag \\
340: X^{i} &=&x^{i}e^{-t} \label{coordinates: planar} \\
341: X^{d} &=&\cosh t-\frac{1}{2}x_{i}x_{i}e^{-t} \notag
342: \end{eqnarray}%
343: with $i=1,\ldots ,d$ the metric take the form%
344: \begin{equation*}
345: ds^{2}=-dt^{2}+e^{-2t}dx_{i}dx_{i} \label{metric: planar}
346: \end{equation*}%
347: In these coordinates the spatial sections have flat \textit{Euclidean}
348: metric.
349:
350: It is useful to introduce conformal coordinates too, defined by the
351: transformation%
352: \begin{equation*}
353: \eta =e^{t}
354: \end{equation*}%
355: The metric is conformally flat and takes the form%
356: \begin{equation*}
357: ds^{2}=\frac{1}{\eta ^{2}}\left( -d\eta ^{2}+dx_{i}dx_{i}\right)
358: \label{metric: Lorentz. conf. flat}
359: \end{equation*}%
360: with $i=1,\ldots ,d$. In the following, unless otherwise stated, we will
361: consider the \textit{Euclidean} section of dS$^{d}$ defined by the
362: analytical continuation%
363: \begin{equation}
364: \eta \rightarrow ix_{0} \label{analitical cont.}
365: \end{equation}%
366: after the transformation (\ref{analitical cont.}) the metric become%
367: \begin{equation}
368: ds^{2}=-\frac{1}{x_{0}^{2}}\left( dx_{0}^{2}+dx_{i}dx_{i}\right)
369: \label{metric: Euclid. conf. flat}
370: \end{equation}%
371: in these coordinates the boundary of dS$^{d}$ $\Sigma $ is given by the
372: submanifold $x_{0}=\epsilon $ where $\epsilon \rightarrow 0$.
373:
374: \subsection{Geodesic Distance\label{Sect.:Geodesic Distance}}
375:
376: The geodesic distance between two points $x$ and $x^{\prime }$ is
377: \begin{equation*}
378: \mu (x,x^{\prime })=\int_{0}^{1}\left[ g_{ab}\dot{x}^{a}(\lambda )\dot{x}%
379: ^{b}(\lambda )\right] ^{\frac{1}{2}}d\lambda
380: ,~~x^{a}(0)=x,~~x^{a}(1)=x^{\prime }
381: \end{equation*}
382: In the following we will often use the new variable%
383: \begin{equation*}
384: z=\cos ^{2}\left( \frac{\mu }{2R}\right) \label{geodes. dist.:z}
385: \end{equation*}%
386: It is possible to show that%
387: \begin{equation*}
388: \cos \left( \frac{\mu (x,x^{\prime })}{R}\right) =\frac{\eta
389: _{ab}X^{a}(x)X^{b}(x^{\prime })}{R^{2}}
390: \end{equation*}%
391: with $X^{a}(x),~X^{b}(x^{\prime })\in \mathcal{M}^{d,1}$ embedding
392: coordinates and $\eta _{ab}=$diag$(-1,1,\ldots ,1)$.
393:
394: Consequently we have%
395: \begin{eqnarray*}
396: z &=&\cos ^{2}\left( \frac{\mu }{2R}\right) \\
397: &=&\frac{1}{2}\left( 1+\cos (\frac{\mu }{R})\right) \\
398: &=&\frac{1}{2}\left( 1+\frac{\eta _{ab}X^{a}(x)X^{b}(x^{\prime })}{R^{2}}%
399: \right)
400: \end{eqnarray*}%
401: In the \textit{Euclidean} conformally flat coordinates (\ref{metric: Euclid.
402: conf. flat}) we have%
403: \begin{equation*}
404: z=-\frac{(x_{0}-y_{0})^{2}+\left( \bar{x}-\bar{y}\right) ^{2}}{x_{0}y_{0}}%
405: =-2+\frac{x_{0}^{2}+y_{0}{}^{2}+(\bar{x}-\bar{y})^{2}}{x_{0}y_{0}}
406: \label{z: euclid. conf. flat coord.}
407: \end{equation*}
408:
409: \subsection{The Cut-off Prescription}
410:
411: Maldacena's prescription defines the boundary correlators by analytic
412: continuation in global time. We have proposed that these formulae should be
413: cut off at a fixed global time $T$. IR divergences will appear as divergent
414: behavior at large $T$. It is most convenient to do calculations in conformal
415: coordinates. Thus we have to understand the effect of a global time cut-off
416: in conformal coordinates.
417:
418: The relation between the two coordinate systems is most simply understood by
419: writing the embedding coordinates in terms of conformal coordinates. The
420: slices of fixed embedding time and global time coincide:
421:
422: %%***************************************************************
423: %Fig 1-2
424: \begin{figure}[tbp]
425: \begin{minipage}[t]{7cm}
426: \begin{center}
427: \includegraphics[width=7.0cm,clip]{PenroseGlobal}
428: \caption[Short caption for figure 2]{ \textbf{Global coordinates.} {\em
429: Foliation of dS with compact spatial sections (spheres).}}
430: \label{PenroseGlobal}
431: \end{center}
432: \end{minipage}
433: \hfill
434: \begin{minipage}[t]{7cm}
435: \begin{center}
436: \includegraphics[width=7.0cm,clip]{PenroseFlat}
437: \caption[Short caption for figure 1]{\textbf{Flat coordinates.} {\em
438: Foliation of dS with flat spatial sections.}} \label{PenroseFlat}
439: \end{center}
440: \end{minipage}
441: \end{figure}
442: %%***************************************************************
443:
444: \begin{equation*}
445: X^{0}=\frac{R}{2}\left( \frac{x^{0}}{R}-\frac{R}{x^{0}}\right) -\frac{%
446: \mathbf{x}^{2}}{2x^{0}} \label{embedconformal}
447: \end{equation*}%
448: At $X^0 = T$, see Fig. \ref{PenroseGlobal} and Fig. \ref{PenroseFlat}. This
449: relation implies a maximal value of $|\mathbf{x}|$ for fixed $x^0$, as well
450: as a maximal value of $x^0$ (which runs between $-\infty$ and $0$ in the
451: conformal coordinate patch). The relation is
452: \begin{equation*}
453: x_{\max }^{2}=-2x^{0}\left( T-x^{0}+\frac{R^{2}}{x^{0}}\right)
454: \end{equation*}%
455: The maximal value of $x^0$ is the point at which $x_{max} = 0$.
456:
457: \begin{equation*}
458: x_{max}^0 \approx - {\frac{R^2 }{T}} \ \ \ T \gg R
459: \end{equation*}
460: The maximal geodesic distance between two points on a give $x^0$ slice is ${%
461: \frac{x_{max} }{x_{max}^0}}$. The slice on which this distance is maximal is
462: given by $x^0_* = - {\frac{2 R^2 }{T}}$. The geodesic distance on this slice
463: is $o(T)$, while the maximum coordinate distance is $o(R)$. IR divergences
464: will come predominantly from slices near this maximal slice.
465:
466: Dirichlet boundary conditions on the $X^0 = T$ surface become \textit{spatial%
467: } Dirichlet boundary conditions on the spatial slices of conformal
468: coordinates. On most of the slice of maximal geodesic size, the Dirichlet
469: propagator will coincide with the usual Euclidean propagator defined by
470: analytic continuation from the entire sphere. Thus, the boundary conditions
471: will not affect the IR divergences.
472:
473: \subsection{Wave Function of the Universe\label{Sect.: Wave Function of the
474: Universe}}
475:
476: We are looking for a gauge invariant definition of the IR\ renormalization
477: of the particle mass. The Wave Function of the Universe (WFU) will provide
478: us with such a definition.
479:
480: The WFU $\Psi \lbrack h_{ij},\phi _{0}]$ was first introduced by Hartle and
481: Hawking in \cite{Hartle:1983ai}. If $I[g,\phi ]$ is the Euclidean action for
482: gravity and a set of fields indicated by $\phi $, the \textit{Euclidean} WFU
483: is defined as the path integral%
484: \begin{equation}
485: \Psi \lbrack h_{ij},\phi _{0}]=\int_{C}[dg][d\phi ]e^{-I[g,\phi ]}~
486: \label{Wave Function Universe: Eucl.}
487: \end{equation}%
488: over a class $C$ of space-times with a compact space-like boundary $\Sigma $
489: on which the induced metric is $h_{ij}$ and over the field configurations $%
490: \phi $ with boundary value $\phi _{0}$. The boundary $\Sigma $ has only one
491: connected component.
492:
493: In the case $\Lambda >0$ we imagine a semiclassical expansion of the
494: integral over Riemannian spaces with the topology of a hemisphere, expanded
495: around the metric on the portion of the round sphere below polar angle $%
496: \theta_0$. We then analytically continue to the future half of Lorentzian dS
497: space. This prescription corresponds to the choice of Euclidean vacuum in de
498: Sitter space.
499:
500: Given the WFU we can define the "boundary two-point function" in the limit
501: where the boundary is taken to $\mathcal{I^+}$%
502: \begin{equation*}
503: \frac{\delta \Psi \lbrack h_{ij},\phi _{0}]}{\delta \phi _{0}(\bar{x})\delta
504: \phi _{0}(\bar{y})}
505: \end{equation*}%
506: Once we expand around dS$^{d}$ we find
507: \begin{equation}
508: \frac{\delta \Psi \lbrack h_{ij},\phi _{0}]}{\delta \phi _{0}(\bar{x})\delta
509: \phi _{0}(\bar{y})}=C_+\frac{1}{(\bar{x}-\bar{y})^{2\Delta_+ }} + C_- \frac{1%
510: }{(\bar{x}-\bar{y})^{2\Delta_- }} \label{Bdry 2-point fnct:
511: dS}
512: \end{equation}%
513: , where $C_{\pm}$ are constants This form is dictated by conformal
514: invariance. If $\lambda $ and $m$ are the coupling and the mass of the field
515: $\phi $, in the classical Lagrangian, then $\Delta $ will be a function of $%
516: \lambda $ and $m$ and will provide a gauge invariant definition of the
517: renormalized mass.
518:
519: The Eq. (\ref{Bdry 2-point fnct: dS}) is the analogue of the boundary
520: correlators defined in the AdS/CFT correspondence%
521: \begin{equation*}
522: Z[\phi _{0}]=\left\langle e^{\int d^{4}x\phi _{0}(x)O(x)}\right\rangle
523: _{CFT}~\ ,~\phi (x_{0}=\epsilon )\sim \phi _{0}
524: \end{equation*}%
525: \begin{equation*}
526: \langle 0|O(\bar{x})O(\bar{y})|0\rangle =\frac{\delta Z}{\delta \phi _{0}(%
527: \bar{x})\delta \phi _{0}(\bar{y})}=\tilde{C}\frac{1}{(\bar{x}-\bar{y})^{2%
528: \tilde{\Delta}}}
529: \end{equation*}
530:
531: There are however, important differences between the two cases. They stem
532: from the fact that the Euclidean section of dS space is a spherical cap and
533: has a conventional Dirichlet problem, different from the singular Dirichlet
534: boundary conditions on the boundary of Euclidean AdS space. There are no
535: large volume divergences in the Euclidean calculation. They appear only
536: after extrapolation to infinite Lorentzian time. As a consequence, the
537: divergent behavior comes as a combination of both powers $\Delta_{\pm}$. For
538: fields corresponding to the principal series of dS representation theory,
539: the real parts of $\Delta_{\pm}$ are equal.
540:
541: The prescription to extract boundary two-point function in dS$^{d}$ given by
542: (\ref{Bdry 2-point fnct: dS}) was first pointed out by Maldacena in \cite%
543: {Maldacena:2002vr} and it is, as explained in this paper, different from the
544: prescription used by Strominger and collaborators in \cite{Strominger:2001pn}%
545: , \cite{Bousso:2001mw}.
546:
547: \subsection{Representations of the dS$^{d}$ Group\label{Representions of dS
548: Group}}
549:
550: The scalar representation of the de Sitter group $SO(1,d)$ are classified
551: according to the mass $m$ in the following series, see \cite{Tolley:2001gg},
552: \cite{Gazeau:1999mi}: the principal series%
553: \begin{equation*}
554: m^{2}\geqslant \left( \frac{d-1}{2R_{dS}}\right) \label{principal
555: series}
556: \end{equation*}%
557: the complementary series
558: \begin{equation*}
559: 0<m^{2}<\left( \frac{d-1}{2R_{dS}}\right) \label{complementary
560: series}
561: \end{equation*}%
562: and the discrete series, whose only case of physical interest is $m^{2}=0$.
563:
564: Under a Wigner-In\"{o}n\"{u} contraction to the Poincare group, only the
565: representations of the principal series contract to representation of the
566: Poincare\ Group.
567:
568: Lowe and G\"{u}ijosa \cite{Guijosa:2003ze} and Lowe \cite{Lowe:2004nw} use
569: the principal series to construct the dS/CFT correspondence. They stress the
570: fact that when one replaces the dS isometry group with a \textit{q}-deformed
571: version, the unitary principal representation deform to a finite dimensional
572: unitary representation of the quantum group\footnote{%
573: The idea that a q-deformed version of the dS group might have finite
574: dimensional unitary representations, resolving the contradiction between dS
575: invariance and a finite number of states, was pointed out to one of the
576: authors (TB) by A. Rajaraman in the fall of 1999. There seemed to be a
577: problem with this idea, because the dS group has no highest weight unitary
578: representations, but Lowe and G\"{u}ijosa made the crucial observation that
579: the cyclic representations of the quantum group (which are not highest
580: weight) converged to the principal unitary series.}.
581:
582: The massive scalar particles in our formulae will always correspond to the
583: principle series representations, so that the boundary dimensions all have
584: the same real part. We will also use a massless, minimally coupled scalar,
585: which is our toy model of the graviton.
586:
587: \section{Scalar Green Functions}
588:
589: In the next few subsections we will derive the scalar Green Functions
590: relevant for our computations and their asymptotic behavior. As explained
591: in the section on the cut-off procedure, we will not impose Dirichlet
592: boundary conditions on the bulk propagators. The IR divergences, which are
593: our principal concern, are not affected by the boundary conditions on the
594: bulk propagator. For a more detailed discussion of dS Green functions, see
595: for example \cite{Allen:1985wd}, \cite{Kirsten:1993ug}.
596:
597: \subsection{Maximally Symmetric Bitensors}
598:
599: The relevant geometric objects in maximally symmetric spaces, like dS, are
600: the geodesic distance $\mu (x,x^{\prime })$ between two points $x$ and $%
601: x^{\prime }$, the unit tangent vectors $n_{\sigma }(x,x^{\prime })$ and $%
602: n_{\sigma ^{\prime }}(x,x^{\prime })$ to the geodesic at $x$ and at $%
603: x^{\prime }$, the vector parallel propagator $g^{\mu }{}_{\nu ^{\prime
604: }}(x,x^{\prime })$ and the spinor parallel propagator $\Lambda ^{\alpha
605: }{}_{\beta ^{\prime }}(x,x^{\prime })$.
606:
607: The geodesic distance is by definition the distance along the geodesic $%
608: x^{a}(\lambda )$ connecting $x$ and $x^{\prime }$%
609: \begin{equation*}
610: \mu (x,x^{\prime })=\int_{0}^{1}\left[ g_{ab}\dot{x}^{a}(\lambda )\dot{x}%
611: ^{b}(\lambda )\right] ^{\frac{1}{2}}d\lambda
612: ,~~x^{a}(0)=x,~~x^{a}(1)=x^{\prime }
613: \end{equation*}%
614: The vectors $n_{\sigma },n_{\sigma ^{\prime }}$ are defined by
615: \begin{equation*}
616: n_{\sigma }=\nabla _{\sigma }\mu (x,x^{\prime })\qquad \mathrm{and}\qquad
617: n_{\sigma ^{\prime }}=\nabla _{\sigma ^{\prime }}\mu (x,x^{\prime })\
618: \end{equation*}%
619: where $\nabla _{\sigma }$ is the covariant derivative. We note that
620: \begin{equation*}
621: n_{\sigma }=-g_{\sigma }{}^{\rho ^{\prime }}n_{\rho ^{\prime }}\
622: \end{equation*}%
623: The vector and spinor parallel propagators are defined by
624: \begin{eqnarray}
625: V^{\mu }(x) &=&g^{\mu }{}_{\nu ^{\prime }}(x,x^{\prime })V^{\nu ^{\prime
626: }}(x^{\prime })\ \label{parallel propag.: vector} \\
627: \psi ^{\alpha }(x) &=&\Lambda ^{\alpha }{}_{\beta ^{\prime }}(x,x^{\prime
628: })\psi ^{\beta ^{\prime }}(x^{\prime }) \label{parallel propag.: spinor}
629: \end{eqnarray}%
630: for every parallel-transported vector $V^{\mu }(x)$ and spinor $\psi
631: ^{\alpha }(x)$, respectively.
632:
633: Tensors that depend on two points $x$ and $x^{\prime }$ on the manifold are
634: called \textit{bitensor. }We will say that a bitensor is \textit{maximally
635: symmetric} if is invariant under any isometry of the manifold. It can be
636: proved that any maximally symmetric bitensor can be expressed as a sum of
637: products of $g^{\mu }{}_{\nu ^{\prime }},~g_{\mu \nu },~g_{\mu ^{\prime }\nu
638: ^{\prime }}{},~\mu ,~n_{\sigma }$ and $n_{\sigma ^{\prime }}$. Furthermore
639: the coefficients of these terms are functions only of the geodesic distance $%
640: \mu (x,x^{\prime })$.
641:
642: The covariant derivatives of the above bitensors are given by
643: \begin{eqnarray}
644: \nabla _{\mu }n_{\nu } &=&A\,(g_{\mu \nu }-n_{\mu }n_{\nu })\ \notag \\
645: \nabla _{\mu ^{\prime }}n_{\nu } &=&C\,(g_{\mu ^{\prime }\nu }+n_{\mu
646: ^{\prime }}n_{\nu })\ \notag \\
647: \nabla _{\mu }g_{\nu \rho ^{\prime }} &=&-(A+C)\,(g_{\mu \nu }n_{\rho
648: ^{\prime }}+g_{\mu \rho ^{\prime }}n_{\nu })\
649: \label{foundam. bitensors: derivatives} \\
650: \nabla _{\mu }\Lambda ^{\alpha }{}_{\beta ^{\prime }} &=&\frac{1}{2}%
651: (A+C)\,[\,(\Gamma _{\mu }\Gamma ^{\nu }n_{\nu }-n_{\mu })\,\Lambda
652: ]^{\,\alpha }{}_{\beta ^{\prime }}\ \notag \\
653: \nabla _{\mu ^{\prime }}\Lambda ^{\alpha }{}_{\beta ^{\prime }} &=&-\frac{1}{%
654: 2}(A+C)\,[\,(\Gamma _{\mu ^{\prime }}\Gamma ^{\nu ^{\prime }}n_{\nu ^{\prime
655: }}-n_{\mu ^{\prime }})\,\Lambda ]^{\,\alpha }{}_{\beta ^{\prime }}\ \notag
656: \end{eqnarray}%
657: where $A$ and $C$ are the following functions of the geodesic distance:
658: \begin{eqnarray}
659: &&\mathrm{for\,\,\,\,\mathbf{R}^{d}:}\qquad A(\mu )=\frac{1}{\mu }\,\qquad
660: C(\mu )=-\frac{1}{\mu }\, \notag \\
661: &&\mathrm{for\,\,\,\,dS~and~AdS:}\qquad A(\mu )=\frac{1}{R}\cot \frac{\mu }{R%
662: }\,\qquad C(\mu )=-\frac{1}{R\sin \left( \frac{\mu }{R}\right) }\
663: \label{table: A, C for R, dS and AdS}
664: \end{eqnarray}%
665: The radius $R$ is real for dS$^{d}$ and it is $R=i\tilde{R}$ with $\tilde{R}$
666: real for AdS$^{d}$. The covariant gamma matrices satisfy the usual relation $%
667: \{\Gamma ^{\mu },\Gamma ^{\nu }\}=2g^{\mu \nu }$.
668:
669: \subsection{Bulk Two-Point Function\label{Scalar Two-Point Function}}
670:
671: In this subsection we will evaluate the scalar two-point function%
672: \begin{equation*}
673: G(x,x^{\prime })=\langle \psi |\phi (x)\phi (x^{\prime })|\psi \rangle
674: \end{equation*}
675:
676: We will assume that the state $|\psi \rangle $ is maximally symmetric, this
677: implies that for spacelike separated points $G(x,x^{\prime })$ depends only
678: on the geodesic distance $\mu (x,x^{\prime })$. For timelike separation the
679: symmetric and Feynman functions also depend only on $\mu $ but the
680: commutator function depend on the time ordering too. Doing an analytical
681: continuation from spacelike separation $\mu ^{2}>0$ to timelike separation $%
682: \mu ^{2}<0$, it is possible to obtain all these two-point functions.
683:
684: We now derive a differential equation for $G(x,x^{\prime })$ . Applying the
685: Laplacian operator to $G(x,x^{\prime })$ we have%
686: \begin{eqnarray*}
687: \square G(\mu ) &=&\nabla ^{\nu }\nabla _{\nu }G(\mu )=\nabla ^{\nu
688: }(G^{\prime }(\mu )n_{\nu }) \\
689: &=&G^{\prime \prime }(\mu )n^{\nu }n_{\nu }+G^{\prime }(\mu )\nabla ^{\nu
690: }n_{\nu } \\
691: &=&G^{\prime \prime }(\mu )+(d-1)A(\mu )G^{\prime }(\mu )
692: \end{eqnarray*}%
693: where we have used the formulae (\ref{table: A, C for R, dS and AdS}) and
694: the notation $G^{\prime }=\frac{dG}{d\mu }$.
695:
696: Using the equation of motion $(\square -m^{2})\phi =0$ we find
697: \begin{equation}
698: G^{\prime \prime }(\mu )+(d-1)A(\mu )G^{\prime }(\mu )-m^{2}G=0
699: \label{Eq for scalar 2-point fnct. in mu}
700: \end{equation}%
701: as long as $x\neq x^{\prime }$.
702:
703: Defining the change of variable%
704: \begin{equation*}
705: z=\cos ^{2}\left( \frac{\mu }{2R}\right)
706: \end{equation*}%
707: the Eq. (\ref{Eq for scalar 2-point fnct. in mu}) for $G$ becomes%
708: \begin{equation}
709: z(1-z)\frac{d^{2}G}{dz^{2}}+[c-(a+b+1)z]\frac{dG}{dz}-abG=0
710: \label{Eq for scalar 2-point fnct. in z}
711: \end{equation}%
712: where we defined%
713: \begin{eqnarray}
714: a &=&\Delta _{+}=\frac{1}{2}\left( d-1+\sqrt{(d-1)^{2}-4m^{2}R^{2}}\right)
715: \label{Delta +} \\
716: b &=&\Delta _{-}=\frac{1}{2}\left( d-1-\sqrt{(d-1)^{2}-4m^{2}R^{2}}\right)
717: \label{Delta -} \\
718: c &=&\frac{1}{2}d \label{c}
719: \end{eqnarray}
720:
721: \subsubsection{De Sitter Space: Massive Scalar\label{SEC: dS GF}}
722:
723: De Sitter space corresponds to choosing $R$ real in the Eq. (\ref{table: A,
724: C for R, dS and AdS}). There are two linearly independent solution $G(z)$ to
725: Eq. (\ref{Eq for scalar 2-point fnct. in z}). Any of the solutions of Eq. (%
726: \ref{Eq for scalar 2-point fnct. in z}) is associated with a particular
727: vacuum $|\psi \rangle $.
728:
729: The Two-point function%
730: \begin{equation*}
731: G_{E}(x,x^{\prime })=\langle E|\phi (x)\phi (x^{\prime })|E\rangle
732: \end{equation*}%
733: associated with the Euclidean vacuum $|E\rangle $ Introduced in Section \ref%
734: {Sect.: Wave Function of the Universe} and defined as analytical
735: continuation from the sphere is given by%
736: \begin{equation}
737: G_{E}(x,x^{\prime })=qF(a,b;c;z) \label{2-point fnct.: scalar, in E
738: vacuum}
739: \end{equation}%
740: where $F(a,b;c;z)$ is the hypergeometric function.
741:
742: The two-point function in the Euclidean vacuum turns out to have the
743: following properties:
744:
745: \begin{enumerate}
746: \item has only one singular point at $\mu (x,x^{\prime })=0$ and therefore
747: regular at $\mu (x,x^{\prime })=\pi R$
748:
749: \item Has the same strength $\mu \rightarrow 0$ singularity as in flat space.
750: \end{enumerate}
751:
752: The constant $q$ in Eq. (\ref{2-point fnct.: scalar, in E vacuum}) is
753: determined by the condition that as $\mu \rightarrow 0$ $G_{E}(x,x^{\prime
754: }) $ has to approach the flat two point function%
755: \begin{equation*}
756: G_{flat}(\mu )\sim \frac{\Gamma \left( \frac{d}{2}\right) }{2(d-2)\pi ^{%
757: \frac{d}{2}}}\mu ^{-d+2},~~\mu \rightarrow 0
758: \end{equation*}%
759: we find%
760: \begin{equation*}
761: q=\frac{\Gamma (a)\Gamma (b)}{\Gamma \left( \frac{d}{2}\right) 2^{d}\pi ^{%
762: \frac{d}{2}}}R^{-d+2}
763: \end{equation*}
764:
765: For the computation it will be useful to derive the asymptotic expansion of $%
766: G(z)$ for $z\rightarrow -\infty $ that correspond to $x_{0}\rightarrow 0$ or
767: $y_{0}\rightarrow 0$.
768:
769: The geodesic distance in conformally flat coordinate was given in Section %
770: \ref{Sect.:Geodesic Distance} and it is
771: \begin{equation*}
772: z=-\frac{(x_{0}-y_{0})^{2}+(\bar{x}-\bar{y})^{2}}{x_{0}y_{0}}
773: \end{equation*}%
774: we have%
775: \begin{equation*}
776: \lim_{\substack{ x_{0}\rightarrow 0 \\ y_{0}\rightarrow 0}}~z\sim -\frac{(%
777: \bar{x}-\bar{y})^{2}}{x_{0}y_{0}}
778: \end{equation*}%
779: so that the asymptotic expansion of $G(z)$ for $z\rightarrow -\infty $ is
780: \begin{equation}
781: \lim_{z\rightarrow \infty }G(z)\sim C_{+}\frac{1}{z^{\Delta _{+}}}+C_{-}%
782: \frac{1}{z^{\Delta _{-}}}=C_{+}\left( \frac{-x_{0}y_{0}}{(\bar{x}-\bar{y}%
783: )^{2}}\right) ^{\Delta _{+}}+C_{-}\left( \frac{-x_{0}y_{0}}{(\bar{x}-\bar{y}%
784: )^{2}}\right) ^{\Delta _{-}} \label{2-point fnct.: scalar, asympt.
785: exp.}
786: \end{equation}%
787: with%
788: \begin{equation*}
789: C_{+}=q~\frac{\Gamma (\frac{d}{2})\Gamma (\Delta _{-}-\Delta _{+})}{\Gamma
790: (\Delta _{-})\Gamma (\frac{d}{2}-\Delta _{+})},~~~C_{-}=q~\frac{\Gamma (%
791: \frac{d}{2})\Gamma (\Delta _{+}-\Delta _{-})}{\Gamma (\Delta _{+})\Gamma (%
792: \frac{d}{2}-\Delta _{-})}
793: \end{equation*}
794:
795: \subsubsection{De Sitter Space: Massless Scalar}
796:
797: The two-point function for a massless minimally coupled scalar field in de
798: Sitter space was studied in \cite{Allen:1987tz}, \cite{Kirsten:1993ug}. They
799: find the following expression for the two-point function%
800: \begin{eqnarray}
801: G_{0}(z) &=&\frac{R^{2}}{192\pi ^{2}m^{2}}+\frac{R}{48\pi ^{2}}\left( \ln
802: (1-z)+\frac{1}{1-z}\right) \label{2-point fnct.:
803: m=0, scalar, in E vacuum}
804: \\
805: &=&C_{0}\left( \ln (1-z)+\frac{1}{1-z}\right) +\tilde{C} \notag
806: \end{eqnarray}
807:
808: We will not need the actual values of the constants $C_{0}$ and $\tilde{C}$
809: in our computation.
810:
811: The asymptotic expansion for $z\rightarrow -\infty $ of the massless
812: two-point function (\ref{2-point fnct.: m=0, scalar, in E vacuum}) is%
813: \begin{equation}
814: G_{0}(z)\sim C_{0}\left( \ln \frac{(\bar{x}-\bar{y})^{2}}{x_{0}y_{0}}+\frac{%
815: x_{0}y_{0}}{(\bar{x}-\bar{y})^{2}}\right)
816: \label{2-point fnct.: m=0,
817: scalar, asympt. exp.}
818: \end{equation}
819:
820: \subsection{Bulk to Boundary Propagators: dS/AdS\label{Sect.: B-B prop.
821: scalar}}
822:
823: The Bulk to Boundary propagator for AdS$^{d}$ were derived by Witten in \cite%
824: {Witten:1998qj}. They obey the equations
825:
826: \begin{equation*}
827: (\square _{x}-m^{2})\tilde{K}(x,\bar{y})=0
828: \end{equation*}%
829: \begin{equation*}
830: \tilde{K}(\bar{x},x_{0};\bar{y})\rightarrow (x_{0})^{((d-1)-\Delta )}\delta
831: ^{d}(\bar{x}-\bar{y}),~\text{for }x_{0}\rightarrow 0
832: \end{equation*}%
833: and their explicit form in the Poincare coordinates in AdS$^{d}$ is%
834: \begin{equation*}
835: \tilde{K}(\bar{x},x_{0};\bar{y})=\frac{\Gamma (\Delta )}{\pi ^{\frac{d-1}{2}%
836: }\Gamma \left( \Delta -\frac{d-1}{2}\right) }\left( \frac{x_{0}}{x_{0}^{2}+(%
837: \bar{x}-\bar{y})^{2}}\right) ^{\Delta }
838: \end{equation*}%
839: with
840: \begin{equation*}
841: \Delta =\Delta _{+}=a=\frac{1}{2}\left( d-1+\sqrt{(d-1)^{2}+4m^{2}\tilde{R}%
842: ^{2}}\right)
843: \end{equation*}%
844: If we consider the conformally flat coordinates (\ref{metric: Euclid. conf.
845: flat}) in dS$^{d}$ the equations defining the Bulk to Boundary propagator
846: become
847:
848: \begin{equation*}
849: \left( \square _{x}+m^{2}\right) K(x,\bar{y})=0
850: \end{equation*}%
851: We impose Dirichlet boundary conditions, $K \rightarrow \delta (x - \bar{y})$
852: as $x$ approaches the boundary of a spherical cap. The cap is then continued
853: to a hemisphere, and analytically continued to $\theta = {\frac{\pi}{2}} + i
854: t$. In our conformal coordinates for the Lorentzian section, $%
855: t\rightarrow\infty$, corresponds to $x^0 \rightarrow 0$. In this limit
856: \begin{equation*}
857: K(\bar{x},x_{0};\bar{y})\rightarrow C_+ (x_{0})^{((d-1)-\Delta_+ )}\delta
858: ^{d}(\bar{x}-\bar{y}) + C_- (x_{0})^{((d-1)-\Delta_- )}\delta ^{d}(\bar{x}-%
859: \bar{y}),~\text{for }x_{0}\rightarrow 0
860: \end{equation*}
861:
862: with%
863: \begin{equation*}
864: \Delta _{\pm}=a=\frac{1}{2}\left( d-1\pm \sqrt{(d-1)^{2}-4m^{2}R^{2}}\right)
865: \end{equation*}
866:
867: \subsection{Boundary Two-point Function: dS/AdS}
868:
869: The boundary two point function for AdS$^{d}$ in the Poincare patch as given
870: for example in \cite{Aharony:1999ti} is
871:
872: \begin{equation*}
873: \langle 0|O(\bar{x})O(\bar{y})|0\rangle =\frac{\delta Z}{\delta \phi (\bar{x}%
874: )\delta \phi (\bar{y})}=C\frac{1}{(\bar{x}-\bar{y})^{2\Delta }}
875: \end{equation*}%
876: with%
877: \begin{equation*}
878: \Delta =\Delta _{+}=a=\frac{1}{2}\left( d-1+\sqrt{(d-1)^{2}+4m^{2}\tilde{R}%
879: ^{2}}\right)
880: \end{equation*}
881:
882: For dS$^{d}$ in the conformally flat coordinates (\ref{metric: Euclid. conf.
883: flat}) we have%
884: \begin{equation*}
885: \frac{\delta \Psi _{0}[h_{ij},\phi _{0}]}{\delta \phi _{0}(\bar{x})\delta
886: \phi _{0}(\bar{y})}=C_+ \frac{1}{(\bar{x}-\bar{y})^{2\Delta _{+}}} + C_-
887: \frac{1}{(\bar{x}-\bar{y})^{2\Delta _{-}}}
888: \end{equation*}%
889: with%
890: \begin{equation*}
891: \Delta _{\pm}=a=\frac{1}{2}\left( d-1 \pm \sqrt{(d-1)^{2}-4m^{2}R^{2}}\right)
892: \end{equation*}
893:
894: \section{General Structure of the Computation\label{Sect.:General Structure
895: of the Computation}}
896:
897: In this section we want to give a general description of the calculation we
898: will perform for three specific models.
899:
900: As we have already discussed in Section \ref{Sect.: Wave Function of the
901: Universe} we are interesting in computing at \textit{1-loop} the Wave
902: Function of the Universe (WFU)%
903: \begin{equation*}
904: \Psi \lbrack h_{ij},\phi _{0}]=\int_{C}[dg][d\phi ]e^{-I[g,\phi ]}~
905: \end{equation*}%
906: for the models described in Section \ref{Sect.: Models}. The \textit{%
907: tree-level} and \textit{1-loop} diagrams are represented respectively in
908: Fig. \ref{treediag} and Fig. \ref{1loopdiag}.
909:
910: Given the WFU we want to find the ``boundary two-point function"%
911: \begin{equation}
912: \frac{\delta \Psi \lbrack h_{ij},\phi _{0}]}{\delta \phi _{0}(\bar{x})\delta
913: \phi _{0}(\bar{y})} \label{Second derivative
914: WFU}
915: \end{equation}%
916: this will provide us with a gauge invariant definition of the renormalized
917: mass.
918:
919: %%***************************************************************
920: %Fig 3-4
921: \begin{figure}[tbp]
922: \begin{minipage}[t]{7cm}
923: \begin{center}
924: \includegraphics[width=7.0cm,clip]{treediag}
925: \caption[Short caption for figure 1]{\textbf{Tree diagram.} {\em This
926: diagram represent the tree-level contribution to the Wave Function of the
927: Universe (WFU). The points $x_1$ and $x_2$ are on the boundary.}}
928: \label{treediag}
929: \end{center}
930: \end{minipage}
931: \hfill
932: \begin{minipage}[t]{7cm}
933: \begin{center}
934: \includegraphics[width=7.0cm,clip]{1loopdiag}
935: \caption[Short caption for figure 2]{\textbf{1-loop diagram.} {\em 1-Loop
936: contribution to the WFU. The points $x_1$ and $x_2$ are on the boundary
937: while $x$ and $y$ are bulk points.}} \label{1loopdiag}
938: \end{center}
939: \end{minipage}
940: \end{figure}
941: %%***************************************************************
942:
943: We consider a general action of the form%
944: \begin{equation*}
945: S=\int d^{d}x\sqrt{g}\left( \phi _{A}\triangle \phi _{A}+\phi _{B}\triangle
946: \phi _{B}+\phi _{C}\triangle \phi _{C}\right) +\lambda \sqrt{g}\phi _{A}\phi
947: _{B}\phi _{C}
948: \end{equation*}%
949: where%
950: \begin{equation*}
951: S_{0}=\int d^{d}x\sqrt{g}\phi _{\alpha }\triangle \phi _{\alpha },~~\alpha
952: =A,B,C
953: \end{equation*}%
954: is the quadratic part of the action i.e.%
955: \begin{equation*}
956: S_{0}=\int d^{d}x\sqrt{g}\frac{1}{2}\left[ (\partial \phi
957: _{A})^{2}+m_{A}^{2}\phi _{A}^{2}\right]
958: \end{equation*}%
959: for a scalar field and%
960: \begin{equation*}
961: S_{0}=S_{M}+S_{\partial M}=\int_{M}d^{d}x\sqrt{g}\bar{\psi}\left(
962: D\!\!\!\!/-m\right) \psi +\int_{\partial M}d^{d}x~\sqrt{h}\bar{\psi}\psi
963: \end{equation*}%
964: for a spinor field.
965:
966: In the WFU we are integrating over fields with the following boundary
967: conditions%
968: \begin{equation*}
969: \phi _{\alpha }|_{\Sigma }=\phi _{\alpha 0},~~\alpha =A,B,C
970: \end{equation*}%
971: where by the symbol $\phi _{\alpha }|_{\Sigma }$ we mean the field evaluated
972: on the boundary of the Euclidean spherical cap. To impose the boundary
973: condition we decompose the field in the following way%
974: \begin{equation*}
975: \phi _{\alpha }=\phi _{\alpha 1}+\phi _{\alpha 2}
976: \end{equation*}%
977: with%
978: \begin{equation*}
979: \phi _{\alpha 1}|_{\Sigma }=\phi _{\alpha 0},~~\phi _{\alpha 2}|_{\Sigma }=0
980: \end{equation*}%
981: The field $\phi _{\alpha 1}$ is the solution of the free wave equation with
982: Dirichlet boundary conditions, and can be written in terms of the
983: appropriate Bulk to Boundary propagator
984: \begin{equation*}
985: \phi _{\alpha 1}=K_{\alpha }\circ \phi _{\alpha 0}=\int_{\Sigma }d\bar{y}%
986: ~K_{\alpha }(\bar{x},x_{0};\bar{y})\phi _{\alpha 0}(\bar{y}),~~\bar{y}\in
987: \Sigma ,~~\alpha =A,B,C
988: \end{equation*}%
989: described in the Sections \ref{Sect.: B-B prop. scalar} and \ref{Sect.: B-B
990: prop. spinor}.
991:
992: To compute the \textit{1-loop }correction to the "boundary two-point
993: function" (\ref{Second derivative WFU}) we have to evaluate the terms in $%
994: \Psi \lbrack h_{ij},\phi _{0}]$ that are quadratic both in $\phi _{0}$ and
995: in the coupling constant $\lambda $. Expanding the path integral we have%
996: \begin{eqnarray*}
997: \Psi &=&\int [d\phi _{A}][d\phi _{B}]~[d\phi _{C}]e^{-S_{0}[\phi _{A},\phi
998: _{B},\phi _{C}]-\int d^{d}x~\sqrt{g(x)}\lambda \phi _{A}\phi _{B}\phi _{C}}
999: \\
1000: &=&\int [d\phi _{A}][d\phi _{B}]~[d\phi _{C}]e^{-S_{0}[\phi _{A},\phi
1001: _{B},\phi _{C}]}\left[ 1-\lambda \int d^{d}x~\sqrt{g(x)}\phi _{A}(x)\phi
1002: _{B}(x)\phi _{C}(x)\right. \\
1003: &&\left. +\frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\sqrt{g(x)}\sqrt{g(y)}%
1004: \phi _{A}(x)\phi _{B}(x)\phi _{C}(x)\phi _{A}(y)\phi _{B}(y)\phi
1005: _{C}(y)+O(\lambda ^{3})\right]
1006: \end{eqnarray*}%
1007: where we indicated with $S_{0}[\phi _{A},\phi _{B},\phi _{C}]$ the quadratic
1008: part of the action.
1009:
1010: The terms quadratic in $\phi _{\alpha 0}~\alpha =A,B,C$ come from the
1011: expansion of the term%
1012: \begin{equation*}
1013: \phi _{A}(x)\phi _{B}(x)\phi _{C}(x)\phi _{A}(y)\phi _{B}(y)\phi _{C}(y)
1014: \end{equation*}%
1015: we have%
1016: \begin{eqnarray*}
1017: &&\phi _{A}(x)\phi _{B}(x)\phi _{C}(x)\phi _{A}(y)\phi _{B}(y)\phi _{C}(y) \\
1018: &=&\phi _{A1}(x)\phi _{A1}(y)\left[ \phi _{B2}(x)\phi _{B2}(y)\phi
1019: _{C2}(x)\phi _{C2}(y)\right] \\
1020: &&+\cdots
1021: \end{eqnarray*}
1022:
1023: We will compute only the correction to the two-point function of the field $%
1024: \phi _{A}$. The part of the path integral relevant to this calculation is
1025: \begin{eqnarray*}
1026: \Psi _{\text{\textit{1-loop}}}^{\text{\textit{A}}}[\phi _{A0}] &=&\int
1027: [d\phi _{B}]~[d\phi _{C}]e^{-\int d^{d}x\sqrt{g}\phi _{B}\triangle \phi
1028: _{B}-\int d^{d}x\sqrt{g}\phi _{C}\triangle \phi _{C}} \\
1029: &&\times \frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\sqrt{g(x)}\sqrt{g(y)}%
1030: \phi _{A1}(x)\phi _{A1}(y)\left[ \phi _{B2}(x)\phi _{B2}(y)\phi _{C2}(x)\phi
1031: _{C2}(y)\right]
1032: \end{eqnarray*}%
1033: The only parts of the fields that fluctuate in the path integral are $\phi
1034: _{\alpha 2}$, in fact $\phi _{\alpha 1}$ is fixed by the boundary
1035: conditions. For this reason the measure of integration is given by%
1036: \begin{equation*}
1037: \lbrack d\phi _{B}]~[d\phi _{C}]=[d\phi _{B2}]~[d\phi _{C2}]
1038: \end{equation*}
1039:
1040: Standard manipulation give us the following expression for the path integral%
1041: \begin{eqnarray*}
1042: \Psi _{\text{\textit{1-loop}}}^{\text{\textit{A}}}[\phi _{A0}] &=&\int
1043: [d\phi _{B2}]~[d\phi _{C2}]e^{-\int d^{d}x\sqrt{g}\phi _{B}\triangle \phi
1044: _{B}-\int d^{d}x\sqrt{g}\phi _{C}\triangle \phi _{C}} \\
1045: &&\times \frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\sqrt{g(x)}\sqrt{g(y)}%
1046: \phi _{A1}(x)\phi _{A1}(y)[\phi _{B2}(x)\phi _{B2}(y)\phi _{C2}(x)\phi
1047: _{C2}(y)] \\
1048: &=&\frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\sqrt{g(x)}\sqrt{g(y)}\phi
1049: _{A1}(x)\phi _{A1}(y)\langle E|\phi _{B2}(x)\phi _{B2}(y)\phi _{C2}(x)\phi
1050: _{C2}(y)|E\rangle \\
1051: &=&\frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\sqrt{g(x)}\sqrt{g(y)}\phi
1052: _{A1}(x)\phi _{A1}(y)\langle E|\phi _{B}(x)\phi _{B}(y)\phi _{C}(x)\phi
1053: _{C}(y)|E\rangle \\
1054: &=&\frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\sqrt{g(x)}\sqrt{g(y)}\phi
1055: _{A1}(x)\phi _{A1}(y)G_{B}(x,y)G_{C}(x,y) \\
1056: &=&\frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\sqrt{g(x)}\sqrt{g(y)}%
1057: \int_{\Sigma }d\bar{x}_{1}~K_{A}(x;\bar{x}_{1})\phi _{A0}(\bar{x}_{1}) \\
1058: &&\times \int_{\Sigma }d\bar{x}_{2}~K_{A}(y;\bar{x}_{2})\phi _{A0}(\bar{x}%
1059: _{2})G_{B}(x,y)G_{C}(x,y)
1060: \end{eqnarray*}
1061:
1062: The boundary two-point function at \textit{1-loop} is given by%
1063: \begin{eqnarray*}
1064: \frac{\delta \Psi ^{\text{\textit{A}}}}{\delta \phi _{A0}(\bar{x}_{1})\delta
1065: \phi _{A0}(\bar{x}_{2})} &=&C\frac{1}{(\bar{x}_{1}-\bar{x}_{2})^{2\Delta }}
1066: \\
1067: &&+\frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\sqrt{g(x)}\sqrt{g(y)}%
1068: ~K_{A}(x;\bar{x}_{1})G_{B}(x,y)G_{C}(x,y)K_{A}(y;\bar{x}_{2})
1069: \end{eqnarray*}
1070:
1071: We have similar expressions for the boundary two-point functions of the
1072: others fields $\phi _{B},~\phi _{C}$.
1073:
1074: \section{Models\label{Sect.: Models}}
1075:
1076: We have computed the \textit{1-loop} boundary two point function for the
1077: following models:
1078:
1079: \textbf{Scalar Fields with Cubic Interaction}%
1080: \begin{equation*}
1081: S=\int d^{d}x\sqrt{g}\frac{1}{2}\left[ (\partial \phi )^{2}+m^{2}\phi
1082: ^{2}+(\partial \phi _{1})^{2}+m^{2}\phi _{1}^{2}+(\partial \phi _{0})^{2}%
1083: \right] +\sqrt{g}\lambda \phi \phi _{0}\phi _{1}
1084: \end{equation*}%
1085: where the field $\phi _{0}$ is massless.
1086:
1087: \textbf{Scalar Fields with Derivative Couplings}
1088:
1089: \begin{equation*}
1090: S=\int d^{d}x\sqrt{g}\frac{1}{2}\left[ (\partial \phi _{A})^{2}+m^{2}\phi
1091: _{A}^{2}+(\partial \phi _{B})^{2}+m^{2}\phi _{B}^{2}+(\partial \phi )^{2}%
1092: \right] +\sqrt{g}\lambda \phi g^{\mu \nu }\partial _{\mu }\phi _{A}\partial
1093: _{\nu }\phi _{B}
1094: \end{equation*}%
1095: where the field $\phi $ is massless.
1096:
1097: \textbf{Spinor Field with Derivative Coupling}
1098:
1099: \begin{equation*}
1100: S_{0}+S_{I}=\int d^{d}x\frac{1}{2}\sqrt{g}(\partial \phi )^{2}+\int_{M}d^{d}x%
1101: \sqrt{g}\bar{\psi}(D\!\!\!\!/-m)\psi +\int_{\partial M}d^{d}x~\sqrt{h}\bar{%
1102: \psi}\psi +\int_{M}d^{d}x~\sqrt{g}\lambda \partial _{a}\phi \bar{\psi}\Gamma
1103: ^{a}\psi
1104: \end{equation*}%
1105: where the field $\phi $ is massless. The surface term for the fermions is
1106: explained in \cite{Arutyunov:1998ve},\cite{Rashkov:1999ji},\cite%
1107: {Henningson:1998cd}.
1108:
1109: We have chosen these models in order to see whether the fact that the
1110: massless boson is derivatively coupled effects the IR divergence, and to
1111: study the effect of fermion chirality. In the conclusions we will discuss
1112: the issues that these results raise for the analogous calculations in
1113: quantum supergravity.
1114:
1115: \section{1-loop Computation: Scalar Fields with Cubic Interaction}
1116:
1117: In this section we will compute the \textit{1-loop} boundary two point
1118: function for the massive field $\phi $ interacting with a massive scalar
1119: field $\phi _{1}$ and a massless scalar field $\phi _{0}$. The lagrangian is%
1120: \begin{equation*}
1121: \mathcal{L}=\sqrt{g}\frac{1}{2}\left[ (\partial \phi )^{2}+m^{2}\phi
1122: ^{2}+(\partial \phi _{1})^{2}+m^{2}\phi _{1}^{2}+(\partial \phi _{0})^{2} %
1123: \right] +\sqrt{g}\lambda \phi \phi _{0}\phi _{1}
1124: \end{equation*}
1125:
1126: The asymptotic expansions of both the bulk and bulk to boundary propagators,
1127: at large Lorentzian time and space-like separation, contain terms with both
1128: powers $(x_0)^{\Delta_{\pm}}$. For the principal series, these powers differ
1129: in the sign of their imaginary part. We have found that the most divergent
1130: terms as $x_0 \rightarrow 0$ come from products of terms from individual
1131: propagators that all have the same power of $x_0$. We call these the \textit{%
1132: pure} terms. Mixed terms have rapidly oscillating phases, which lead to more
1133: convergent integrals. We will find that in this model the pure terms look
1134: like the tree level results, but with a divergent correction to the mass.
1135: The mixed terms are sub-leading, and do not have the same form as the tree
1136: level result. We will explicitly show only our results for the pure terms.
1137:
1138: As explained in Section \ref{Sect.:General Structure of the Computation} the
1139: \textit{1-loop} correction to the boundary two-point function%
1140: \begin{equation*}
1141: \frac{\delta \Psi _{\text{\textit{1-loop}}}}{\delta \phi _{0}(\bar{x}%
1142: _{1})\delta \phi _{0}(\bar{x}_{2})}=G_{\text{\textit{1-loop}}}(\bar{x}_{1},%
1143: \bar{x}_{2})
1144: \end{equation*}
1145:
1146: is given by%
1147: \begin{eqnarray*}
1148: G_{\text{\textit{1-loop}}}(\bar{x}_{1},\bar{x}_{2}) &=&\frac{\lambda ^{2}}{2}%
1149: \int d^{d}x\int d^{d}y~\sqrt{g(x)}\sqrt{g(y)}~K(x;\bar{x}%
1150: _{1})G_{1}(x,y)G_{0}(x,y)K(y;\bar{x}_{2}) \\
1151: &=&\frac{\lambda ^{2}}{2}\int d^{d-1}\bar{x}\int d^{d-1}\bar{y}\int
1152: dx_{0}\int dy_{0}~\frac{1}{x_{0}^{d}}\frac{1}{y_{0}^{d}}~K(x;\bar{x}%
1153: _{1})G_{1}(x,y)G_{0}(x,y)K(y;\bar{x}_{2})
1154: \end{eqnarray*}%
1155: In principle, the bulk propagators in these equations should satisfy
1156: (vanishing) Dirichlet boundary conditions at a fixed global time, $T$. We
1157: have seen that in conformal coordinates this corresponds to an $x^{0}$
1158: dependent Dirichlet boundary condition on a sphere in $\mathbf{x}$ space, as
1159: well as an upper cut-off $x_{max}^{0}\sim -R^{2}/T$. The IR divergences will
1160: come from the regions of maximal spatial geodesic size, and, because of the
1161: Dirichlet boundary conditions, from regions where the two integrated bulk
1162: points are far from the spatial boundary sphere. Thus considering only the
1163: leading IR divergent part of the answer, we can use the usual Euclidean
1164: vacuum Green's function (without Dirichlet boundary conditions) and
1165: approximate it by its asymptotic form at large geodesic distance:%
1166: \begin{eqnarray*}
1167: G_{\text{\textit{1-loop}}}^{\text{\textit{IR}}}(\bar{x}_{1},\bar{x}_{2})
1168: &\sim &\frac{\lambda ^{2}}{2}\int d^{d-1}\bar{x}\int d^{d-1}\bar{y}\int
1169: dx_{0}\int dy_{0}~\frac{1}{x_{0}^{d}}\frac{1}{y_{0}^{d}}%
1170: ~(x_{0}y_{0})^{((d-1)-\Delta _{\pm })} \\
1171: &&\times \delta ^{d-1}(\bar{x}-\bar{x}_{1})G_{1}(x,y)G_{0}(x,y)\delta ^{d-1}(%
1172: \bar{y}-\bar{x}_{2}) \\
1173: &=&\frac{\lambda ^{2}}{2}\int dx_{0}\int dy_{0}~\frac{1}{x_{0}^{d}}\frac{1}{%
1174: y_{0}^{d}}~(x_{0}y_{0})^{((d-1)-\Delta _{\pm })}G_{0}(\bar{x}_{1},x_{0};\bar{%
1175: x}_{2},y_{0})G_{1}(\bar{x}_{1},x_{0};\bar{x}_{2},y_{0}) \\
1176: &\sim &\frac{\lambda ^{2}}{2}\int_{\alpha }^{\epsilon }dx_{0}\int_{\beta
1177: }^{\epsilon }dy_{0}~\frac{1}{x_{0}}\frac{1}{y_{0}}~C_{0}C_{-}\ln \left(
1178: \frac{(\bar{x}-\bar{y})^{2}}{x_{0}y_{0}}\right) \left( \frac{1}{(\bar{x}-%
1179: \bar{y})^{2}}\right) ^{\Delta _{\pm }}
1180: \end{eqnarray*}%
1181: Here we used the fact that bulk to boundary propagators satisfy
1182:
1183: \begin{equation*}
1184: K(\bar{x},x_{0};\bar{y})\rightarrow C_{+}(x_{0})^{((d-1)-\Delta _{+})}\delta
1185: ^{d}(\bar{x}-\bar{y})+C_{-}(x_{0})^{((d-1)-\Delta _{-})}\delta ^{d}(\bar{x}-%
1186: \bar{y}),~\text{for }x_{0}\rightarrow 0
1187: \end{equation*}%
1188: explained in Section \ref{Sect.: B-B prop. scalar} and the asymptotic
1189: expansion (\ref{2-point fnct.: scalar, asympt. exp.}), (\ref{2-point fnct.:
1190: m=0, scalar, asympt. exp.}) for the bulk two-point functions\footnote{%
1191: In tree level calculations involving two bulk to boundary propagators, only
1192: one of them can be replaced by a $\delta $ function, since the other ends up
1193: evaluated at separated points. The powers of $x^{0}$ that would set it equal
1194: to zero are part of the renormalization factor that defines the limiting
1195: boundary two point function. In our calculation, both bulk to boundary
1196: propagators are legitimately replaced by $\delta $ functions.} .
1197:
1198: Integrating in $x_{0}$ and $y_{0}$ and keeping the leading part in $\epsilon
1199: \rightarrow 0$ we find%
1200: \begin{eqnarray}
1201: G_{\text{\textit{1-loop}}}^{\text{\textit{IR}}}(\bar{x}_{1},\bar{x}_{2})
1202: &\sim &\frac{\lambda ^{2}}{2}\frac{1}{(\bar{x}_{1}-\bar{x}_{2})^{2\Delta
1203: _{\pm }}}\times \left( \ln \left( \frac{(\bar{x}_{1}-\bar{x}_{2})^{2}}{%
1204: \epsilon }\right) \right) ^{3} \notag \\
1205: &&+\text{Subleading terms in }\epsilon \notag
1206: \end{eqnarray}
1207:
1208: \section{1-loop Computation: Scalar Fields with Derivative Coupling}
1209:
1210: In this section we will compute the \textit{1-loop} boundary two points
1211: function for the massive scalar field $\phi $ derivatively coupled to a
1212: massless scalar field $\phi _{A}$ and a massive scalar field $\phi _{B}$.
1213: The action is
1214:
1215: \begin{equation*}
1216: S=\int d^{d}x\sqrt{g}\frac{1}{2}\left[ (\partial \phi )^{2}+m^{2}\phi
1217: ^{2}+(\partial \phi _{B})^{2}+m^{2}\phi _{B}^{2}+(\partial \phi _{A})^{2}%
1218: \right] +\sqrt{g}\lambda \phi g^{\mu \nu }\partial _{\mu }\phi _{A}\partial
1219: _{\nu }\phi _{B}
1220: \end{equation*}
1221:
1222: Following the general lines of the computation done in Section \ref%
1223: {Sect.:General Structure of the Computation} we find for the \textit{1-loop}
1224: WFU
1225:
1226: \begin{eqnarray*}
1227: \Psi _{\text{\textit{1-loop}}} &=&\int [d\phi _{B2}]~[d\phi _{C2}]e^{-\int
1228: d^{d}x\sqrt{g}\frac{1}{2}\left[ (\partial \phi )^{2}+m^{2}\phi
1229: ^{2}+(\partial \phi _{B})^{2}+m^{2}\phi _{B}^{2}+(\partial \phi _{A})^{2}%
1230: \right] } \\
1231: &&\times \frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\sqrt{g(x)}\sqrt{g(y)}
1232: \\
1233: &&\times (\phi (x)g^{\mu \nu }(x)\partial _{\mu }\phi _{A}(x)\partial _{\nu
1234: }\phi _{B}(x))(\phi (y)g^{\rho \lambda }(y)\partial _{\rho }\phi
1235: _{A}(y)\partial _{\lambda }\phi _{B}(y)) \\
1236: &=&\frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\sqrt{g(x)}\sqrt{g(y)}\phi
1237: _{1}(x)\phi _{1}(y) \\
1238: &&\times g^{\mu \nu }(x)g^{\rho \lambda }(y)\partial _{\mu }^{x}\partial
1239: _{\rho }^{y}G_{A}(x,y)\partial _{\nu }^{x}\partial _{\lambda }^{y}G_{B}(x,y)
1240: \\
1241: &=&\frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\sqrt{g(x)}\sqrt{g(y)}%
1242: \int_{\Sigma }d\bar{x}_{1}~K_{A}(x;\bar{x}_{1})\phi _{0}(\bar{x}%
1243: _{1})\int_{\Sigma }d\bar{x}_{2}~K_{A}(y;\bar{x}_{2})\phi _{0}(\bar{x}_{2}) \\
1244: &&\times g^{\mu \nu }(x)g^{\rho \lambda }(y)\partial _{\mu }^{x}\partial
1245: _{\rho }^{y}G_{A}(x,y)\partial _{\nu }^{x}\partial _{\lambda }^{y}G_{B}(x,y)
1246: \end{eqnarray*}
1247:
1248: The \textit{1-loop} two point function is%
1249: \begin{equation*}
1250: \frac{\delta \Psi _{\text{\textit{1-loop}}}}{\delta \phi _{0}(\bar{x}%
1251: _{1})\delta \phi _{0}(\bar{x}_{2})}=G_{\text{\textit{1-loop}}}(\bar{x}_{1},%
1252: \bar{x}_{2})
1253: \end{equation*}
1254:
1255: Considering only the leading IR divergent part we have%
1256: \begin{eqnarray*}
1257: G_{\text{\textit{1-loop}}}^{\text{\textit{IR}}}(\bar{x}_{1},\bar{x}_{2})
1258: &\sim &\frac{\lambda ^{2}}{2}\int dx_{0}\int dy_{0}~\frac{1}{x_{0}^{d}}\frac{%
1259: 1}{y_{0}^{d}}~(x_{0}y_{0})^{((d-1)-\Delta _{\pm })} \\
1260: &&\times g^{\mu \nu }(x)g^{\rho \lambda }(y)\partial _{\mu }^{x}\partial
1261: _{\rho }^{y}C_{0}C_{-}\ln \left( \frac{(\bar{x}_{1}-\bar{x}_{2})^{2}}{%
1262: x_{0}y_{0}}\right) \partial _{\nu }^{x}\partial _{\lambda }^{y}\left( \frac{%
1263: x_{0}y_{0}}{(\bar{x}_{1}-\bar{x}_{2})^{2}}\right) ^{\Delta _{\pm }} \\
1264: &=&\frac{\lambda ^{2}}{2}\int dx_{0}\int dy_{0}~\frac{1}{x_{0}^{d}}\frac{1}{%
1265: y_{0}^{d}}~x_{0}^{2}y_{0}^{2}(x_{0}y_{0})^{((d-1)-\Delta _{\pm })}\partial
1266: _{\mu }^{x}\partial _{\rho }^{y}C_{0}C_{-} \\
1267: &&\times \ln \left( \frac{(\bar{x}_{1}-\bar{x}_{2})^{2}}{x_{0}y_{0}}\right)
1268: \partial _{\mu }^{x}\partial _{\rho }^{y}\left( \frac{x_{0}y_{0}}{(\bar{x}%
1269: _{1}-\bar{x}_{2})^{2}}\right) ^{\Delta _{\pm }} \\
1270: &=&\frac{\lambda ^{2}}{2}\int dx_{0}\int dy_{0}~\frac{1}{x_{0}^{d}}\frac{1}{%
1271: y_{0}^{d}}~x_{0}^{2}y_{0}^{2}(x_{0}y_{0})^{((d-1)-\Delta _{\pm })}\partial
1272: _{i}^{x}\partial _{j}^{y}C_{0}C_{-} \\
1273: &&\times \ln \left( (\bar{x}_{1}-\bar{x}_{2})^{2}\right) \partial
1274: _{i}^{x}\partial _{j}^{y}\left( \frac{x_{0}y_{0}}{(\bar{x}_{1}-\bar{x}%
1275: _{2})^{2}}\right) ^{\Delta _{\pm }} \\
1276: &=&\frac{\lambda ^{2}}{2}\int dx_{0}\int dy_{0}~x_{0}^{1}y_{0}^{1}(-4\Delta
1277: (3+2\Delta ))C_{0}C_{-}\left( \frac{1}{(\bar{x}_{1}-\bar{x}_{2})^{2}}\right)
1278: ^{2+\Delta _{\pm }}
1279: \end{eqnarray*}%
1280: where we used the bulk to boundary propagators property explained in Section %
1281: \ref{Sect.: B-B prop. scalar} and the asymptotic expansion (\ref{2-point
1282: fnct.: scalar, asympt. exp.}), (\ref{2-point fnct.: m=0, scalar, asympt.
1283: exp.}) for the bulk two-point functions. Furthermore we used the fact that%
1284: \begin{equation*}
1285: \partial _{0}^{x}\partial _{0}^{y}\left( \ln \frac{(\bar{x}-\bar{y})^{2}}{%
1286: x_{0}y_{0}}\right) =0,~~\partial _{0}^{x}\partial _{j}^{y}\left( \ln \frac{(%
1287: \bar{x}-\bar{y})^{2}}{x_{0}y_{0}}\right) =0
1288: \end{equation*}%
1289: with $i,j=1,\ldots ,d$ .
1290:
1291: Doing the integrals and keeping the leading parts in $\epsilon \rightarrow 0$
1292: we find%
1293: \begin{eqnarray}
1294: G_{\text{\textit{1-loop}}}^{\text{\textit{IR}}}(\bar{x}_{1},\bar{x}_{2})
1295: &\sim &(\epsilon )^{4}\left( \frac{1}{(\bar{x}_{1}-\bar{x}_{2})^{2}}\right)
1296: ^{2+\Delta _{-}}
1297: \label{Bdry 2-point fnct 1-loop: dS, 3 scalar deriv. coupl.} \\
1298: &&+\text{Subleading terms in }\epsilon \notag
1299: \end{eqnarray}
1300:
1301: \section{1-loop Computation: Spinor Field with Derivative Coupling}
1302:
1303: In this last section we will evaluate the \textit{1-loop} boundary two-point
1304: function for a spinor field $\psi $ derivatively coupled to a massless
1305: scalar field $\phi $. The action in the tangent frame is
1306:
1307: \begin{equation*}
1308: S_{0}+S_{I}=\int_{M}d^{d}x\frac{1}{2}\sqrt{g}(\partial \phi
1309: )^{2}+\int_{M}d^{d}x\sqrt{g}\bar{\psi}(D\!\!\!\!/-m)\psi +\int_{\partial
1310: M}d^{d}x~\sqrt{h}\bar{\psi}\psi +\int_{M}d^{d}x~\lambda \sqrt{g}\partial
1311: _{a}\phi \bar{\psi}\Gamma ^{a}\psi
1312: \end{equation*}%
1313: The surface term for the fermions is explained in \cite{Arutyunov:1998ve},%
1314: \cite{Rashkov:1999ji},\cite{Henningson:1998cd}.
1315:
1316: More specifically we are using the metric%
1317: \begin{equation*}
1318: ds^{2}=-\frac{1}{x_{0}^{2}}(dx^{0}dx^{0}+d\bar{x}\cdot d\bar{x})=-\frac{1}{%
1319: x_{0}^{2}}(dx^{0}dx^{0}+dx_{i}dx_{i})
1320: \end{equation*}
1321:
1322: and the vielbein $e_{\mu }^{a},~a=0,\ldots ,d-1$ such that $g_{\mu \nu }=$ $%
1323: e_{\mu }^{a}e_{\nu }^{b}\eta _{ab}$. The explicit form of the vielbein and
1324: is inverse is%
1325: \begin{eqnarray*}
1326: e_{\mu }^{a} &=&\frac{1}{x_{0}}\delta _{\mu }^{a} \\
1327: e_{a}^{\mu } &=&x_{0}\delta _{a}^{\mu }
1328: \end{eqnarray*}%
1329: the spin connection has the form%
1330: \begin{equation*}
1331: \omega _{i}^{0j}=\omega _{i}^{j0}=\frac{1}{x_{0}}\delta _{i}^{j}
1332: \end{equation*}%
1333: and all other component vanishing. The Dirac operator is given by%
1334: \begin{equation*}
1335: D\!\!\!\!/=e_{a}^{\mu }(\partial _{\mu }+\frac{1}{2}\omega _{\mu
1336: }^{bc}\Sigma _{bc})=x_{0}\Gamma ^{0}\partial _{0}+x_{0}\bar{\Gamma}\cdot
1337: \nabla -\frac{d-1}{2}\Gamma ^{0}
1338: \end{equation*}%
1339: where $\Gamma ^{a}=(\Gamma ^{0},\Gamma ^{i})=(\Gamma ^{0},\bar{\Gamma})$
1340: satisfy $\{\Gamma ^{a},\Gamma ^{b}\}=2\eta ^{ab}$ and $\partial _{\mu
1341: }=(\partial _{0},\partial _{i})=(\partial _{0},\nabla )$.
1342:
1343: The explicit form of the interacting term is
1344: \begin{equation*}
1345: \mathcal{L}_{I}=\lambda \sqrt{g}\partial _{a}\phi \bar{\psi}\Gamma ^{a}\psi
1346: =\lambda \sqrt{g}e_{a}^{\mu }\partial _{\mu }\phi \bar{\psi}\Gamma ^{a}\psi
1347: =\lambda \sqrt{g}x_{0}\delta _{a}^{\mu }\partial _{\mu }\phi \bar{\psi}%
1348: \Gamma ^{a}\psi
1349: \end{equation*}
1350:
1351: Again following the same reasoning of Section \ref{Sect.:General Structure
1352: of the Computation} we find for the \textit{1-loop} WFU%
1353: \begin{eqnarray*}
1354: \Psi _{\text{\textit{1-loop}}} &=&\int [d\psi ]~[d\bar{\psi}]e^{-\left(
1355: \int_{M}d^{d}x\frac{1}{2}\sqrt{g}(\partial \phi )^{2}+\int_{M}d^{d}x\sqrt{g}%
1356: \bar{\psi}(D\!\!\!\!/-m)\psi +\int_{\partial M}d^{d}x~\sqrt{h}\bar{\psi}\psi
1357: +\int_{M}d^{d}x~\lambda \sqrt{g}\partial _{a}\phi \bar{\psi}\Gamma ^{a}\psi
1358: \right) } \\
1359: &=&\int [d\psi ]~[d\bar{\psi}]e^{-S_{0}}\int d^{d}x\int d^{d}y~\sqrt{g(x)}%
1360: \sqrt{g(y)} \\
1361: &&\times \frac{\lambda ^{2}}{2}\left( \partial _{a}\phi (x)\bar{\psi}%
1362: (x)\Gamma ^{a}\psi (x)\right) \left( \partial _{b}\phi (y)\bar{\psi}%
1363: (y)\Gamma ^{b}\psi (y)\right) \\
1364: &=&\frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\sqrt{g(x)}\sqrt{g(y)}\bar{%
1365: \psi}_{1}(x)\langle E|\partial _{a}\phi (x)\Gamma ^{a}\psi (x)\partial
1366: _{b}\phi (y)\bar{\psi}(y)\Gamma ^{b}|E\rangle \psi _{1}(y) \\
1367: &=&\frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\sqrt{g(x)}\sqrt{g(y)}\bar{%
1368: \psi}_{1}(x)\Gamma ^{a}S(x,y)\Gamma ^{b}\partial _{a}^{x}\partial
1369: _{b}^{y}G_{0}(x,y)\psi _{1}(y) \\
1370: &=&\frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\sqrt{g(x)}\sqrt{g(y)}\int
1371: d^{d-1}\bar{x}~\bar{\psi}_{0}(\bar{x})K(y,\bar{x})\Gamma ^{a}S(x,y)\Gamma
1372: ^{b}\partial _{a}^{x}\partial _{b}^{y}G_{0}(x,y) \\
1373: &&\times \int d^{d-1}\bar{x}~K(x,\bar{x})\psi _{0}(\bar{x})
1374: \end{eqnarray*}
1375:
1376: taking the limit $x_{0}\rightarrow 0,~y_{0}\rightarrow 0$ we find the
1377: leading IR part of $\Psi _{\text{\textit{1-loop}}}$
1378: \begin{eqnarray*}
1379: \Psi _{\text{\textit{1-loop}}}^{\text{\textit{IR}}} &\sim &\frac{\lambda ^{2}%
1380: }{2}\int d^{d}x\int d^{d}y~\frac{1}{x_{0}^{d}}\frac{1}{y_{0}^{d}}\bar{\psi}%
1381: _{0+}(\bar{x}_{1})\Gamma ^{a}S(x,y)\Gamma ^{b}\partial _{a}^{x}\partial
1382: _{b}^{y}G_{0}(x,y)\psi _{0-}(\bar{x}_{2}) \\
1383: &\sim &\frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\frac{1}{x_{0}^{d}}\frac{%
1384: 1}{y_{0}^{d}} \\
1385: &&\times \bar{\psi}_{0+}(\bar{x}_{1})\Gamma ^{a}C_{-}C_{0}\left( \frac{%
1386: x_{0}y_{0}}{(\bar{x}-\bar{y})^{2}}\right) ^{\Delta _{-}}\frac{\bar{\Gamma}%
1387: \cdot (\bar{x}-\bar{y})}{\left\vert \bar{x}-\bar{y}\right\vert }\Gamma
1388: ^{b}\partial _{a}^{x}\partial _{b}^{y}\ln \left( \frac{(\bar{x}-\bar{y})^{2}%
1389: }{x_{0}y_{0}}\right) \psi _{0-}(\bar{x}_{2}) \\
1390: &=&\frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\frac{1}{x_{0}^{d}}\frac{1}{%
1391: y_{0}^{d}}(x_{0}y_{0})^{\Delta _{1}+1} \\
1392: &&\times \bar{\psi}_{0+}(\bar{x}_{1})\Gamma ^{a}C_{-}C_{0}\left( \frac{%
1393: x_{0}y_{0}}{(\bar{x}-\bar{y})^{2}}\right) ^{\Delta _{-}}\frac{\bar{\Gamma}%
1394: \cdot (\bar{x}-\bar{y})}{\left\vert \bar{x}-\bar{y}\right\vert }\Gamma
1395: ^{b}\delta _{a}^{\mu }\partial _{\mu }^{x}\delta _{b}^{\nu }\partial _{\nu
1396: }^{y}\ln \left( \frac{(\bar{x}-\bar{y})^{2}}{x_{0}y_{0}}\right) \psi _{0-}(%
1397: \bar{x}_{2}) \\
1398: &=&\frac{\lambda ^{2}}{2}\int d^{d}x\int d^{d}y~\frac{1}{x_{0}^{d}}\frac{1}{%
1399: y_{0}^{d}}(x_{0}y_{0})^{\Delta _{-}+1} \\
1400: &&\times \bar{\psi}_{0+}(\bar{x}_{1})\Gamma ^{a}C_{-}C_{0}\left( \frac{%
1401: x_{0}y_{0}}{(\bar{x}-\bar{y})^{2}}\right) ^{\Delta _{-}}\frac{\bar{\Gamma}%
1402: \cdot (\bar{x}-\bar{y})}{\left\vert \bar{x}-\bar{y}\right\vert }\Gamma
1403: ^{b}\delta _{a}^{i}\partial _{i}^{x}\delta _{b}^{j}\partial _{j}^{y}\ln
1404: \left( (\bar{x}-\bar{y})^{2}\right) \psi _{0-}(\bar{x}_{2}) \\
1405: &=&\frac{\lambda ^{2}}{2}C_{-}C_{0}\int d^{d}x\int d^{d}y~\frac{1}{x_{0}^{d}}%
1406: \frac{1}{y_{0}^{d}}(x_{0}y_{0})^{\Delta _{-}+1} \\
1407: &&\times \bar{\psi}_{0+}(\bar{x}_{1})\Gamma ^{i}\left( \frac{1}{(\bar{x}-%
1408: \bar{y})^{2}}\right) ^{\Delta _{-}}\frac{\Gamma ^{k}(\bar{x}-\bar{y})_{k}}{%
1409: \left\vert \bar{x}-\bar{y}\right\vert }\Gamma ^{j}\partial _{i}^{x}\partial
1410: _{j}^{y}\ln \left( (\bar{x}-\bar{y})^{2}\right) \psi _{0-}(\bar{x}_{2})
1411: \end{eqnarray*}%
1412: where we used the bulk to boundary propagators property%
1413: \begin{equation*}
1414: \lim_{x_{0}\rightarrow 0}(x_{0})^{-\frac{d}{2}+m}\psi (x)=-c\psi _{0-}(\bar{x%
1415: })
1416: \end{equation*}%
1417: \begin{equation*}
1418: \lim_{x_{0}\rightarrow 0}(x_{0})^{-\frac{d}{2}+m}\bar{\psi}(x)=c\bar{\psi}%
1419: _{0+}(\bar{x})
1420: \end{equation*}%
1421: explained in Appendix \ref{Sect.: B-B prop. spinor} and the asymptotic
1422: expansion (\ref{2-point fnct.: spinor, asympt. exp.}), (\ref{2-point fnct.:
1423: m=0, scalar, asympt. exp.}) for the bulk two-point functions. As in the
1424: previous section we noticed that%
1425: \begin{equation*}
1426: \partial _{0}^{x}\partial _{0}^{y}\ln \left( \frac{(\bar{x}-\bar{y})^{2}}{%
1427: x_{0}y_{0}}\right) =0,~~\partial _{0}^{x}\partial _{j}^{y}\ln \left( \frac{(%
1428: \bar{x}-\bar{y})^{2}}{x_{0}y_{0}}\right) =0,~~i,j=1,\ldots ,d
1429: \end{equation*}
1430:
1431: The boundary two-point function at 1-loop is%
1432: \begin{eqnarray*}
1433: G_{\text{\textit{1-loop}}}(\bar{x}_{1},\bar{x}_{2}) &=&\frac{\delta \Psi _{%
1434: \text{\textit{1-loop}}}^{\text{\textit{IR}}}}{\delta \bar{\psi}_{0+}(\bar{x}%
1435: _{1})\delta \psi _{0-}(\bar{x}_{2})} \\
1436: &=&\frac{\lambda ^{2}}{2}C_{0}C_{-}\int_{\alpha }^{\epsilon
1437: }d^{0}x\int_{\beta }^{\epsilon }d^{0}y~(x_{0}y_{0})^{(\Delta _{-}+1-d)}~ \\
1438: &&\times \Gamma ^{i}\left( \frac{1}{(\bar{x}_{1}-\bar{x}_{2})^{2}}\right)
1439: ^{\Delta _{-}}\frac{\Gamma ^{k}(\bar{x}_{1}-\bar{x}_{2})_{k}}{\left\vert
1440: \bar{x}_{1}-\bar{x}_{2}\right\vert }\Gamma ^{j}\partial _{i}^{x}\partial
1441: _{j}^{y}\ln \left( (\bar{x}_{1}-\bar{x}_{2})^{2}\right)
1442: \end{eqnarray*}%
1443: doing the integrals and keeping the leading terms in $\epsilon \rightarrow 0$
1444: we find%
1445: \begin{eqnarray}
1446: G_{\text{\textit{1-loop}}}(\bar{x}_{1},\bar{x}_{2}) &\sim &\lambda
1447: ^{2}\epsilon ^{2(\Delta _{-}-d+2)}
1448: \label{Bdry 2-point fnct 1-loop: dS, 1 scalar + 1 fermion} \\
1449: &&\times \Gamma ^{i}\Gamma ^{k}\Gamma ^{j}\left( \frac{1}{(\bar{x}_{1}-\bar{x%
1450: }_{2})^{2}}\right) ^{\Delta _{-}}\frac{(\bar{x}_{1}-\bar{x}_{2})_{k}}{%
1451: \left\vert \bar{x}_{1}-\bar{x}_{2}\right\vert }\partial _{i}^{x}\partial
1452: _{j}^{y}\ln \left( (\bar{x}_{1}-\bar{x}_{2})^{2}\right) \notag \\
1453: &&+\text{Subleading terms in }\epsilon \notag
1454: \end{eqnarray}
1455:
1456: \section{Analysis Divergences}
1457:
1458: \subsection{Three Massive Scalar Fields with Cubic Interaction in dS$^{d}$}
1459:
1460: \subsubsection{Leading Terms}
1461:
1462: We didn't perform explicitly the computation in this case but it is easy to
1463: see that the leading IR divergent term (which is not in fact divergent in
1464: this case) in the boundary two-point function has the following form up to a
1465: constant
1466:
1467: \begin{eqnarray*}
1468: G_{\text{\textit{1-loop}}}^{\text{\textit{IR}}}(\bar{x}_{1},\bar{x}_{2})
1469: &\sim &\frac{\lambda ^{2}}{2}\epsilon ^{2\Delta _{2}}\frac{1}{(\bar{x}_{1}-%
1470: \bar{x}_{2})^{2\Delta _{2}}}\frac{1}{(\bar{x}_{1}-\bar{x}_{2})^{2\Delta _{1}}%
1471: } \\
1472: &&+\text{Subleading terms in }\epsilon
1473: \end{eqnarray*}%
1474: where $\Delta _{2},~\Delta _{1}$ correspond respectively to the fields $\phi
1475: _{1}$ and $\phi _{2}$. In this expression we have kept only pure terms.
1476: Other terms are no more divergent than these.
1477:
1478: The leading IR term in $G_{\text{\textit{1-loop}}}^{\text{\textit{IR}}}(\bar{%
1479: x}_{1},\bar{x}_{2})$ is proportional to%
1480: \begin{equation*}
1481: \epsilon ^{2\Delta _{2}}
1482: \end{equation*}
1483:
1484: We have%
1485: \begin{equation*}
1486: \Delta _{\pm }^{i}=\frac{1}{2}\left( d-1\pm \sqrt{%
1487: (d-1)^{2}-4m_{i}^{2}R_{dS}^{2}}\right) =\frac{1}{2}(d-1)\left( 1\pm \sqrt{%
1488: (1-\alpha _{i})}\right)
1489: \end{equation*}%
1490: with%
1491: \begin{equation*}
1492: \alpha _{i}=\left( \frac{2m_{i}R_{dS}}{d-1}\right) ^{2}
1493: \end{equation*}%
1494: So we immediately see that $G_{\text{\textit{1-loop}}}^{\text{\textit{IR}}}(%
1495: \bar{x}_{1},\bar{x}_{2})$ is IR convergent for every $\alpha _{i}$ i.e. both
1496: for the complementary and principal series, see Section \ref{Representions
1497: of dS Group}.
1498:
1499: This computation shows that in the case of massive fields there is no IR
1500: divergence in the boundary two point function. This is in accord with naive
1501: expectations.
1502:
1503: \subsection{Two Massive and One Massless field in dS$^{d}$}
1504:
1505: The leading IR term in $G_{\text{\textit{1-loop}}}(\bar{x}_{1},\bar{x}_{2})$
1506: is proportional to%
1507: \begin{equation*}
1508: (\log \epsilon )^{3}
1509: \end{equation*}
1510:
1511: So in this case $G_{\text{\textit{1-loop}}}(\bar{x}_{1},\bar{x}_{2})$ is IR
1512: divergent.
1513:
1514: The analysis of divergences in the remaining case (\ref{Bdry 2-point fnct
1515: 1-loop: dS, 3 scalar deriv. coupl.}), (\ref{Bdry 2-point fnct 1-loop: dS, 1
1516: scalar + 1 fermion}) is very similar and we will not repeat it. We want only
1517: to remark that these cases are not IR\ divergent, due to the presence of
1518: derivative couplings, as can be seen inspecting the power dependence of the $%
1519: \epsilon $ cutoff.
1520:
1521: \section{The Meaning of the Divergences}
1522:
1523: To understand the meaning of the divergences we have found, we compare our
1524: expressions to those obtained by perturbing the free massive theory by a
1525: term ${\frac{1}{2}} \delta m^2 \phi^2$. That computation gives
1526:
1527: \begin{equation*}
1528: \delta m^{2}\int dx_{0}~\frac{1}{x_{0}^{d}}\int d^{d-1}\bar{x}~K(x_{0},\bar{x%
1529: };\bar{x}_{b})K(x_{0},\bar{x};\bar{y}_{b})
1530: \end{equation*}%
1531: where $K$ is the massive bulk to boundary propagator. The IR divergent
1532: contribution to this integral comes from $x_{0}\sim 0$, where we can
1533: substitute one of the propagators by $K(x_{0},\bar{x};\bar{x}_{b})\sim
1534: (x_{0})^{d-1-\Delta }\delta (\bar{x}-\bar{x}_{b})$. The result is%
1535: \begin{equation*}
1536: \delta m^{2}\int dx_{0}~\frac{1}{x_{0}}\left\vert \bar{x}_{b}-\bar{y}%
1537: _{b}\right\vert ^{-2\Delta }
1538: \end{equation*}%
1539: It is important to note that this expression for the perturbed two point
1540: function could be derived explicitly from the expression of the two point
1541: function as an integral over the boundary. One simply uses Green's theorem
1542: and a perturbative analysis of the Klein-Gordon equation. The same statement
1543: would \textit{not} be true in AdS/CFT. In that context, the Euclidean
1544: boundary conditions depend on $\delta m^{2}$, and so the straightforward
1545: perturbative analysis of the path integral misses a term coming from the
1546: perturbation of the boundary conditions. It turns out that the missing term
1547: is sub-leading if the boundary operator is irrelevant, but is the dominant
1548: term if it is marginal or relevant.
1549:
1550: By contrast, in the one loop computation with massless fields and
1551: non-derivative coupling, we obtained the IR divergent part%
1552: \begin{equation*}
1553: \int dx_{0}\int dy_{0}~\frac{1}{x_{0}^{d}}\frac{1}{y_{0}^{d}}%
1554: (x_{0}y_{0})^{d-1-\Delta }\left( \frac{x_{0}y_{0}}{\left\vert \bar{x}_{b}-%
1555: \bar{y}_{b}\right\vert ^{2}}\right) ^{\Delta }\left( \ln x_{0}+\ln
1556: y_{0}\right)
1557: \end{equation*}%
1558: The first term after the integration measure comes from the two bulk to
1559: boundary propagators, which we have approximated by their small $x_{0}$
1560: limits. This enabled us to do the two spatial integrals using the $\delta $
1561: functions. The first term in square brackets is the asymptotic form of the
1562: massive bulk propagator, while the second is that of the massless
1563: propagator. We note that if we had instead exchanged a massive field from
1564: the principle series in the loop, or if the massless scalar had derivative
1565: couplings, this last factor would have been a positive power of $x_{0}$ and
1566: all the integrals in the loop diagram would have been convergent. This means
1567: that for a purely massive theory the IR region of coordinate space does not
1568: contribute to the mass renormalization at all\footnote{%
1569: We would get contributions from the region where the two bulk points in the
1570: diagram were close together, corresponding to the usual UV mass
1571: renormalization.}. The value of the mass renormalization following from
1572: exchange of a minimal massless scalar, with soft couplings is thus%
1573: \begin{equation*}
1574: \delta m^{2}\propto \int dx_{0}~\frac{1}{x_{0}}\ln x_{0}\sim \ln ^{2}T\sim
1575: \ln ^{2}\Lambda
1576: \end{equation*}%
1577: The last equality reflects our prejudice that the IR cutoff should be
1578: determined in terms of the c.c., by the requirement of finite entropy.
1579:
1580: We note that minimally coupled scalars would generally arise as
1581: Nambu-Goldstone bosons and would be derivatively coupled. Our calculation
1582: shows that one would not expect IR mass divergences in models with NG
1583: bosons. However, we believe that there are indications that gravity has IR
1584: divergence problems comparable to those of minimally coupled massless bosons
1585: with soft couplings. Thus, the divergence we have uncovered reflects our
1586: best guess at the behavior of perturbative quantum gravity in dS space.
1587:
1588: \section{Generalization to a Model with Gravity}
1589:
1590: The simplest generalization of the calculations we have done is to a model
1591: of gravity interacting with a massive scalar in a dS background. The
1592: Lagrangian is
1593: \begin{equation*}
1594: \mathcal{L}=\sqrt{\left\vert g\right\vert }\left[ M_{P}^{2}R-\left( g^{\mu
1595: \nu }\partial _{\mu }\phi \partial _{\nu }\phi +m^{2}\phi ^{2}\right) \right]
1596: \end{equation*}%
1597: As always in perturbative quantum gravity calculations must be done in a
1598: fixed gauge. We first studied this problem in the gauge for fluctuations
1599: around the dS metric defined by
1600: \begin{equation*}
1601: h_{\mu \nu }={\frac{1}{d}}g_{\mu \nu }h+H_{\mu \nu }
1602: \end{equation*}%
1603: \begin{equation*}
1604: g^{\mu \nu }H_{\mu \nu }=0=D^{\mu }H_{\mu \nu }
1605: \end{equation*}%
1606: $g_{\mu \nu }$ is the background dS metric, and $D^{\mu }$ its Christoffel
1607: connection. In this gauge, the Lagrangian for $h$ is that of a scalar field
1608: with tachyonic mass, while the components of $H_{\mu \nu }$ satisfy a
1609: massive Klein-Gordon equation. One might think that the IR divergences at
1610: one loop arise only from the exchange of $h$\footnote{%
1611: In this gauge, ghosts couple only to gravitons and so there are no ghost
1612: contributions to the one loop boundary two point function of the massive
1613: scalar.}. If this were the case, the calculation would be a simple
1614: generalization of our non-derivative trilinear scalar interaction, with the
1615: massless field replaced by a tachyon.
1616:
1617: The result of this computation is disastrous and confusing. The IR
1618: divergence is power law rather than logarithmic (relative to the tree level
1619: calculation). Furthermore the power of $\left\vert \bar{x}_{b}-\bar{y}%
1620: _{b}\right\vert $ differs from the tree level power, so we cannot interpret
1621: the effect as a mass renormalization. If this result were valid one would be
1622: led to the conclusion that the dS/CFT correlation functions simply did not
1623: exist, even in perturbation theory, and the divergence could not be
1624: explained as a divergent mass renormalization.
1625:
1626: We gained insight by viewing the transverse gauge as the $\alpha \rightarrow
1627: 0$ limit of the one parameter family of gauge fixing Lagrangians
1628: \begin{equation*}
1629: \delta \mathcal{L=}\frac{1}{2\alpha }\left( D^{\mu }H_{\mu \nu }+2b\alpha
1630: \partial _{\nu }h\right) ^{2}
1631: \end{equation*}%
1632: The coefficient $b$ is chosen to cancel the mixing between $H_{\mu \nu }$
1633: and $h$ in the classical Lichnerowicz Lagrangian for fluctuations around dS
1634: space. In this class of gauges, it is easy to see that the tachyonic mass,
1635: as well as the overall normalization of the $h$ propagator, is $\alpha $
1636: dependent. The same is therefore true of the power of $T$ and of $\left\vert
1637: \bar{x}_{b}-\bar{y}_{b}\right\vert $ in the the IR divergent part of the $h$
1638: exchange graph.
1639:
1640: Thus, either this contribution is canceled by $H_{\mu\nu}$ exchange, or the
1641: answer is not gauge invariant. Formal arguments using graphical Ward
1642: identities seem to suggest that the boundary two point function is indeed $%
1643: \alpha$ independent. Thus, we expect the power law IR divergences to cancel
1644: at this order. This suggests the possibility that logarithmic divergences,
1645: which come from the behavior of the transverse, traceless part of the
1646: graviton propagator, may not cancel. Gravitational theories would then
1647: exhibit the same sort of IR divergences as our toy model. Of course, we
1648: really need to do a careful computation in order to verify gauge invariance
1649: of the results. We plan to return to this in a future publication. See \cite%
1650: {Henningson:1998cd} and references therein.
1651:
1652: \appendix
1653:
1654: \section{Comparison with AdS}
1655:
1656: In this appendix we record comparisons of our computation of three massive
1657: scalars, with an analogous computation of AdS space. The purpose of this is
1658: to verify that there is no analogue of the divergences we have found, even
1659: when one of the scalars is massless. The essential reason for this
1660: difference is that the bulk AdS propagator is constructed only from
1661: normalizable modes. By contrast, in dS space the Euclidean propagator
1662: contains both solutions of the homogeneous wave equation at large proper
1663: distance.
1664:
1665: \subsection{Three Scalar Fields AdS}
1666:
1667: For comparison we will describe the case of three massive scalar fields with
1668: cubic interaction in AdS.
1669:
1670: As before it is easy to see that in AdS\ the part of $G_{\text{\textit{1-loop%
1671: }}}^{\text{\textit{IR}}}(\bar{x}_{1},\bar{x}_{2})$ that is dependent on $%
1672: \epsilon $ is proportional to%
1673: \begin{equation*}
1674: \epsilon ^{2\Delta _{+}}
1675: \end{equation*}%
1676: In AdS we consider only one type of modes%
1677: \begin{equation*}
1678: \Delta =\Delta _{+}=\frac{1}{2}\left( d-1+\sqrt{%
1679: (d-1)^{2}+4m_{i}^{2}R_{AdS}^{2}}\right) =\frac{1}{2}(d-1)\left( 1+\sqrt{%
1680: (1+\alpha _{i})}\right)
1681: \end{equation*}%
1682: with%
1683: \begin{equation*}
1684: \alpha _{i}=\left( \frac{2m_{i}R_{AdS}}{d-1}\right) ^{2}
1685: \end{equation*}%
1686: so%
1687: \begin{equation*}
1688: \Delta _{+}>0,~\forall ~\alpha _{i}
1689: \end{equation*}%
1690: and $G_{\text{\textit{1-loop}}}(\bar{x}_{1},\bar{x}_{2})$ is IR convergent
1691: for every $\alpha _{i}$ even when $m_{i}$ is zero.
1692:
1693: \subsubsection{Anti de Sitter:\ Scalar Propagator}
1694:
1695: The two-point function for a scalar field of mass $m$ in AdS$^{d}$ has been
1696: derived for example in \cite{Allen:1985wd}. They find%
1697: \begin{eqnarray}
1698: G(z) &=&rz^{-a}F(a,a-c+1;a-b+1;z^{-1}) \label{2-point fnct.: AdS scalar} \\
1699: r &=&\frac{\Gamma (a)\Gamma (a-c+1)}{\Gamma (a-b+1)\pi ^{\frac{d}{2}}2^{d}}%
1700: R^{2-d} \notag
1701: \end{eqnarray}%
1702: with $a,~b,~c$ given respectively by (\ref{Delta +}), (\ref{Delta -}), (\ref%
1703: {c}) and where for AdS$^{d}$ we have $R=i\tilde{R},~\tilde{R}\in
1704: %TCIMACRO{\U{211d} }%
1705: %BeginExpansion
1706: \mathbb{R}
1707: %EndExpansion
1708: $.
1709:
1710: The asymptotic expansion $z\rightarrow \infty $ of (\ref{2-point fnct.: AdS
1711: scalar}) is
1712:
1713: \begin{equation*}
1714: F(a,a-c+1;a-b+1;z^{-1})\rightarrow 1
1715: \end{equation*}
1716:
1717: \begin{equation*}
1718: \lim_{z\rightarrow \infty }G(z)\sim rz^{-\Delta }
1719: \end{equation*}%
1720: with
1721:
1722: \begin{equation*}
1723: \Delta =\Delta _{+}=a=\frac{1}{2}\left( d-1+\sqrt{(d-1)^{2}+4m^{2}\tilde{R}%
1724: ^{2}}\right)
1725: \end{equation*}
1726:
1727: \section{Spinor Green Functions\label{Spinor Two-point Function}}
1728:
1729: Here we record the spinor Green Functions needed for the computations and
1730: their asymptotic behavior. For a more exhaustive discussion see for example
1731: \cite{Allen:1985wd}, \cite{Anguelova:2003kf}, \cite{Muck:1999mh}.
1732:
1733: \subsection{Spinor Parallel Propagator}
1734:
1735: In this section we will derive a differential equation for the spinor
1736: parallel propagator $\Lambda (x^{\prime },x)_{\;\beta }^{\alpha ^{\prime }}$
1737: (\ref{parallel propag.: spinor}) whose action on a spinor is
1738:
1739: \begin{equation*}
1740: \psi {^{\prime }}(x^{\prime })^{\alpha ^{\prime }}=\Lambda (x^{\prime
1741: },x)_{\;\beta }^{\alpha ^{\prime }}\psi (x)^{\beta } \label{def big lambda}
1742: \end{equation*}%
1743: this equation for $\Lambda (x^{\prime },x)_{\;\beta }^{\alpha ^{\prime }}$
1744: will be a fundamental ingredient in the derivation of the spinor Green
1745: function $S(x,x^{\prime })$ .
1746:
1747: $\Lambda (x^{\prime },x)$ satisfy the following properties
1748: \begin{subequations}
1749: \begin{align}
1750: n^{\mu }\nabla _{\mu }\Lambda (x,x^{\prime })& =0 \label{property
1751: lambda 1}
1752: \\
1753: \Lambda (x^{\prime },x)& =[\Lambda (x,x^{\prime })]^{-1}
1754: \label{property lambda 2} \\
1755: \Gamma ^{\nu ^{\prime }}(x^{\prime })& =\Lambda (x^{\prime },x)\Gamma ^{\mu
1756: }(x)\Lambda (x,x^{\prime })g_{\mu }^{\nu ^{\prime }}(x^{\prime },x)
1757: \label{property lambda 3}
1758: \end{align}%
1759: (\ref{property lambda 1}) follows from the definition of parallel transport
1760: of a spinor along a curve,\ (\ref{property lambda 2}) derive from the fact
1761: that the $\Lambda (x^{\prime },x)$ form a group and (\ref{property lambda 3}%
1762: )\ indicate how to parallel transport the gamma matrices.
1763:
1764: Manipulating the previous equations we obtain
1765: \end{subequations}
1766: \begin{equation}
1767: \nabla _{\mu }\Lambda (x,x^{\prime })=\frac{1}{2}(A+C)\left( \Gamma _{\mu
1768: }\Gamma ^{\nu }n_{\nu }-n_{\mu }\right) \Lambda (x,x^{\prime })
1769: \label{d lambda}
1770: \end{equation}%
1771: and%
1772: \begin{equation*}
1773: \nabla _{\mu ^{\prime }}\Lambda (x,x^{\prime })=-\frac{1}{2}(A+C)\Lambda
1774: (x,x^{\prime })\left( \Gamma _{\mu ^{\prime }}\Gamma ^{\nu ^{\prime }}n_{\nu
1775: ^{\prime }}-n_{\mu ^{\prime }}\right) \label{d lamda prime}
1776: \end{equation*}
1777:
1778: \subsection{Bulk Two-Point Function}
1779:
1780: The spinor Green $S(x,x^{\prime })$ function is defined by the equation
1781:
1782: \begin{equation}
1783: \left[ (\slashD-m)S(x,x^{\prime })\right] _{\;\beta ^{\prime }}^{\alpha }=%
1784: \frac{\delta (x-x^{\prime })}{\sqrt{g(x)}}\delta _{\beta ^{\prime }}^{\alpha
1785: } \label{definition spinor GF}
1786: \end{equation}%
1787: The most general form for $S(x,x^{\prime })$ is%
1788: \begin{equation}
1789: S(x,x^{\prime })=\left[ \alpha (\mu )+\beta (\mu )n_{\nu }\Gamma ^{\nu }%
1790: \right] \Lambda (x,x^{\prime }) \label{Spinor GF general form}
1791: \end{equation}%
1792: with $\alpha (\mu ),~\beta (\mu )$ functions only of the geodesic distance.
1793:
1794: Substituting (\ref{Spinor GF general form}) into (\ref{definition spinor GF}%
1795: ) and using (\ref{d lambda})\ we obtain two differential equations for $%
1796: \alpha (\mu )$ and $\ \beta (\mu )$
1797: \begin{align}
1798: \beta ^{\prime }+\frac{1}{2}(d-1)(A-C)\beta -m\alpha & =\frac{\delta
1799: (x-x^{\prime })}{\sqrt{g(x)}} \label{Eqs for alpha beta 1} \\
1800: \alpha ^{\prime }+\frac{1}{2}(d-1)(A+C)\alpha -m\beta & =0,
1801: \label{Eqs for alpha beta 2}
1802: \end{align}%
1803: Combining (\ref{Eqs for alpha beta 1}) and (\ref{Eqs for alpha beta 2}) we
1804: find the following differential equation for $\alpha (\mu )$%
1805: \begin{equation}
1806: \alpha ^{\prime \prime }+(d-1)A\alpha ^{\prime }-\frac{1}{2}%
1807: (d-1)C(A+C)\alpha -\left[ \frac{(d-1)^{2}}{4R^{2}}+m^{2}\right] \alpha =m%
1808: \frac{\delta (x-x^{\prime })}{\sqrt{g(x)}} \label{Eq for alpha}
1809: \end{equation}
1810:
1811: \subsubsection{De Sitter Space: Massive Spinor}
1812:
1813: To derive $S(x,x^{\prime })$ in dS$^{d}$ space we perform the change of
1814: variables%
1815: \begin{eqnarray*}
1816: z &=&\cos ^{2}\frac{\mu }{2R} \label{change variables alpha z} \\
1817: \alpha (z) &=&\sqrt{z}\gamma (z) \notag
1818: \end{eqnarray*}%
1819: the Eq. (\ref{Eq for alpha}) become
1820: \begin{subequations}
1821: \label{gammaeq}
1822: \begin{gather}
1823: H(a,b;c;z)\gamma (z)=0 \label{Hypergeom. eq for gamma} \\
1824: H(a,b;c;z)=z(1-z)\frac{d^{2}}{dz^{2}}+[c-(a+b+1)z]\frac{d}{dz}-ab \notag
1825: \end{gather}%
1826: with
1827: \end{subequations}
1828: \begin{equation*}
1829: a=\frac{d}{2}-i|m|R,\quad b=\frac{d}{2}+i|m|R,\quad c=\frac{d}{2}+1
1830: \label{a, b, c for spinor}
1831: \end{equation*}
1832:
1833: As explained in Section \ref{SEC: dS GF},\ the solution corresponding to the
1834: \textit{Euclidean vacuum} is the one that is singular only at $z=1$ i.e.
1835: \begin{equation*}
1836: \gamma (z)=\lambda \,\mathrm{F}(a,b;c;z)=\lambda \,\mathrm{F}%
1837: (d/2-i|m|R,d/2+i|m|R;d/2+1;z) \label{gamma for S}
1838: \end{equation*}%
1839: \begin{equation*}
1840: \alpha (z)=\lambda \sqrt{z}\,\mathrm{F}(d/2-i|m|R,d/2+i|m|R;d/2+1;z)
1841: \label{alpha for
1842: S}
1843: \end{equation*}%
1844: The constant $\lambda $ is derived by the requirement that (\ref{Spinor GF
1845: general form}) has the same behavior of the flat spinor Green function for $%
1846: R\rightarrow \infty $. We have
1847: \begin{equation*}
1848: \lambda =-m\frac{\Gamma (d/2-i|m|R)\Gamma (d/2+i|m|R)}{\Gamma (d/2+1)\pi
1849: ^{d/2}2^{d}}R^{2-d} \label{constant lambda for S}
1850: \end{equation*}%
1851: Finally $\beta (z)$ is determined by the Eq. (\ref{Eqs for alpha beta 2})
1852: \begin{align}
1853: \beta (z)& =-\frac{1}{m}\left[ \frac{1}{R}\sqrt{z(1-z)}\frac{d}{dz}+\frac{d-1%
1854: }{2R}\sqrt{\frac{1-z}{z}}\right] \alpha (z) \label{beta for S} \\
1855: & =-\frac{\lambda }{mR}\sqrt{1-z}\left[ z\,\mathrm{F}%
1856: (d/2+1-i|m|R,d/2+1+i|m|R;d/2+2;z)\phantom{\frac{n}2}\right. \notag \\
1857: & \quad +\left. \frac{d}{2}\,\mathrm{F}(d/2-i|m|R,d/2+i|m|R;d/2+1;z)\right]
1858: \notag
1859: \end{align}
1860: The asymptotic $z\rightarrow -\infty $ expansion for the spinor two-point
1861: function is found to be
1862: \begin{equation}
1863: \lim_{\substack{ x_{0}\rightarrow 0 \\ ~y_{0}\rightarrow 0}}S(x,y)=\left(
1864: \left( C_{+}\frac{-x_{0}y_{0}}{(\overline{x}-\overline{y})^{2}}\right)
1865: ^{\Delta _{+}}+C_{-}\left( \frac{-x_{0}y_{0}}{(\overline{x}-\overline{y})^{2}%
1866: }\right) ^{\Delta _{-}}\right) \frac{\bar{\Gamma}\cdot \left( \overline{x}-%
1867: \overline{y}\right) }{\left\vert \overline{x}-\overline{y}\right\vert }
1868: \label{2-point fnct.:
1869: spinor, asympt. exp.}
1870: \end{equation}%
1871: with%
1872: \begin{eqnarray*}
1873: \Delta _{+} &=&\frac{d-1}{2}+im \\
1874: \Delta _{-} &=&\frac{d-1}{2}-im
1875: \end{eqnarray*}
1876:
1877: \subsection{Bulk to Boundary Propagators: dS/AdS\label{Sect.: B-B prop.
1878: spinor}}
1879:
1880: The complete expression for the spinor Bulk to Boundary propagators:
1881:
1882: \begin{equation}
1883: \psi _{1}(x)=\int d^{d-1}\bar{x}~K(x,\bar{x})\psi _{0}(\bar{x})
1884: \label{Bulk to Bndry: spinor psi}
1885: \end{equation}%
1886: \begin{equation}
1887: \bar{\psi}_{1}(x)=\int d^{d-1}\bar{x}~\bar{\psi}_{0}(\bar{x})K(x,\bar{x}%
1888: )\psi _{0}(\bar{x}) \label{Bulk to Bndry: spinor psi bar}
1889: \end{equation}%
1890: has been given for example in \cite{Henningson:1998cd}.
1891:
1892: For our purposes we will need only the asymptotic expansion $%
1893: x_{0}\rightarrow 0,~y_{0}\rightarrow 0$ for the propagators (\ref{Bulk to
1894: Bndry: spinor psi}), (\ref{Bulk to Bndry: spinor psi bar}), we have
1895: \begin{equation}
1896: \lim_{x_{0}\rightarrow 0}(x_{0})^{-\frac{d}{2}+m}\left( -\frac{1}{c}\right)
1897: \psi (x)=\psi _{0-}(\bar{x})-\frac{1}{c}\int d^{d-1}\bar{y}~\left\vert \bar{x%
1898: }-\bar{y}\right\vert ^{-d-1+2m}(\bar{x}-\bar{y})\cdot \bar{\Gamma}\psi _{0+}(%
1899: \bar{y}) \label{Bulk to Bndry: spinor psi, asympt. exp.}
1900: \end{equation}%
1901: \begin{equation}
1902: \lim_{x_{0}\rightarrow 0}(x_{0})^{-\frac{d}{2}+m}\left( \frac{1}{c}\right)
1903: \bar{\psi}(x)=\bar{\psi}_{0+}(\bar{x})+\frac{1}{c}\int d^{d-1}\bar{y}~\bar{%
1904: \psi}_{0-}(\bar{y})(\bar{x}-\bar{y})\cdot \bar{\Gamma}\left\vert \bar{x}-%
1905: \bar{y}\right\vert ^{-d-1+2m}
1906: \label{Bulk to Bndry: spinor psi
1907: bar, asympt. exp.}
1908: \end{equation}
1909:
1910: where the constant is $c=\pi ^{d/2}\Gamma (m+\frac{1}{2})/\Gamma (m+\frac{d+1%
1911: }{2})$. And we have used the following decomposition for the fields
1912:
1913: \begin{eqnarray*}
1914: \psi _{0}(\bar{x}) &=&\psi _{0+}(\bar{x})+\psi _{0-}(\bar{x}) \\
1915: \bar{\psi}_{0}(\bar{x}) &=&\bar{\psi}_{0+}(\bar{x})+\bar{\psi}_{0-}(\bar{x})
1916: \end{eqnarray*}%
1917: with%
1918: \begin{eqnarray*}
1919: \Gamma ^{0}\psi _{\pm }(\bar{x}) &=&\pm \psi _{\pm }(\bar{x}) \\
1920: \bar{\psi}_{\pm }(\bar{x})\Gamma ^{0} &=&\pm \bar{\psi}_{\pm }(\bar{x})
1921: \end{eqnarray*}%
1922: For the right-hand side of (\ref{Bulk to Bndry: spinor psi, asympt. exp.}), (%
1923: \ref{Bulk to Bndry: spinor psi bar, asympt. exp.}) to be integrable, with
1924: respect to the measure $d^{d-1}\bar{y}$ on the boundary $\Sigma $ we have to
1925: impose the conditions%
1926: \begin{eqnarray*}
1927: \psi _{+}(\bar{y}) &=&0 \\
1928: \bar{\psi}_{-}(\bar{y}) &=&0
1929: \end{eqnarray*}
1930:
1931: %------------------------------ ACKNOWLEDGMENTS ----------------------------
1932:
1933: \acknowledgments
1934:
1935: TB and LM would like to acknowledge conversations with J. Maldacena about
1936: his approach to the dS/CFT correspondence. Their work was supported in part
1937: by the DOE under grant DE-FG03-92ER40689. The work of WF was supported in
1938: part by the NSF under grant 0071512.
1939:
1940: \bibliographystyle{utcaps}
1941: \bibliography{acompat,gravitino_mass_bibtex}
1942:
1943: \end{document}
1944: