Inflation and String Theory: Daniel Baumann and Liam Mcallister
Inflation and String Theory: Daniel Baumann and Liam Mcallister
Inflation and String Theory: Daniel Baumann and Liam Mcallister
F
D.A.M.T.P., Cambridge University, Cambridge, CB3 0WA, UK
Department of Physics, Cornell University, Ithaca, NY 14853, USA
Abstract
We review cosmological inflation and its realization in quantum field theory
and in string theory. This material is a portion of a book, also entitled
Inflation and String Theory, to be published by Cambridge University Press.
Contents
Preface 7
2
Contents 3
2.1.2 Naturalness 58
2.1.3 Symmetries 62
2.1.4 Gravity 64
2.1.5 Time-Dependence 68
2.2 Effective Theories of Inflation 69
2.2.1 Slow-Roll: Dynamics and Perturbations 70
2.2.2 Slow-Roll: Selected Models 72
2.2.3 Non-Slow-Roll: K-Inflation 79
2.2.4 Inflation in Effective Field Theory 80
2.3 Ultraviolet Sensitivity 81
2.3.1 Eta Problem I: Radiative Corrections 81
2.3.2 Eta Problem II: Higher-Dimension Operators 83
2.3.3 Gravity Waves and Super-Planckian Fields 84
2.3.4 Non-Gaussianity 92
References 362
Preface
7
8 Preface
1
Pure Einstein gravity is free of one-loop divergences, but diverges at two loops. Gravi-
tational theories including matter fields typically diverge at one loop, except in super-
symmetric cases [21].
10 Preface
The boldest hope for the use of string theory in cosmology is that string
theory will open entirely new dynamical realms that cannot be described
in any effective quantum field theory with a finite number of fields, and
the resulting cosmic histories will avoid or overcome the limitations of con-
temporary models. While this enticing prospect has inspired work in string
cosmology for more than two decades, in our opinion string theory is not
yet understood at the level required for such a dramatic step. Even the low-
energy effective actions governing the interactions of massless string states
in non-supersymmetric vacua are not adequately characterized at present,
while computing dynamics driven by the full tower of massive strings is a
distant dream. Fundamental advances in understanding time-dependent so-
lutions of string theory with string scale curvatures will be required if we are
to move outside the aegis of the effective theory for the massless modes. In
this book we will restrict our attention to conservative applications of string
theory to the study of inflation: we will survey the substantial literature in
which string theory underpins or informs inflationary effective theories, but
does not replace them outright.
2
The appendices will appear in the final version of the book [49], but are omitted from
the arXiv version.
Preface 13
14
Notation and Conventions 15
H
2 , ,
H H
where overdots stand for derivatives with respect to physical time t. The
potential slow-roll parameters are
2 2
V0 00
Mpl 2 V
, Mpl ,
2 V V
where primes are derivatives with respect to the inflaton , and V () is the
potential energy density.
We define the string length and the string mass, respectively, as
1
`2s 0 , Ms2 ,
0
where 0 is the Regge slope. Beware of factors of 2 in alternative defini-
tions of these quantities in the literature. The ten-dimensional gravitational
coupling is
22 = (2)7 (0 )4 .
1
Inflation: Theory and Observations
so that the distance |x| (the comoving distance) that a particle can travel
between times 1 and 2 = 1 + is simply |x| = , for any a( ). In
the standard Big Bang cosmology, the expansion at early times is driven
by the energy density of radiation, and by tracing the evolution backward
16
1.1 Horizon Problem 17
one finds that a 0 at sufficiently early times, and the metric becomes
singular at this point. We choose coordinates so that the initial singularity
is at t = 0. At some time t > 0, the maximum comoving distance a particle
can have traversed since the initial singularity (a.k.a the particle horizon)
is given by
dt0
Z t Z a
d ln a 1 da
= 0 = , where H . (1.3)
0 a(t ) 0 aH a dt
During the standard Big Bang evolution, a < 0 and the comoving Hubble
1 1
radius (aH) = (a) grows with time. The integral in (1.3) is therefore
dominated by the contributions from late times. This leads to the so-called
horizon problem. The amount of conformal time that elapses between the
singularity and the formation of the cosmic microwave background (an event
known as recombination) is much smaller than the conformal time between
recombination and today (see fig. 1.1). Quantitatively, one finds that points
50 1000 10 3 1 0 1 3 10 1000
now 1.0
0.8
ne
40
co
conformal time [Gyr]
0.6
ht
lig
scale factor
0.4
30
0.2
20
ere
0.1
ph
le s
10
bb
Hu
0.01
CMB 0.001
0
-40 -20 0 20 40
comoving distance [Glyr]
Fig. 1.1. Spacetime diagram illustrating the horizon problem in comoving coor-
dinates (figure adapted from [52]). The dotted vertical lines correspond to the
worldlines of comoving objects. We are the central worldline. The current red-
shifts of the comoving galaxies are labeled on each worldline. All events that we
currently observe are on our past light cone. The intersection of our past light
cone with the spacelike slice labeled CMB corresponds to two opposite points on
the CMB surface of last-scattering. The past light cones of these points, shaded
gray, do not overlap, so the points appear never to have been in causal contact.
in the CMB that are separated by more than one degree were never in
causal contact, according to the standard cosmology: their past light cones
do not overlap before the spacetime is terminated by the initial singularity.
Yet their temperatures are observed to be the same to one part in 104 .
Moreover, the observed temperature fluctuations are correlated on what
seem to be superhorizon scales at recombination. Not only must we explain
18 1 Inflation: Theory and Observations
why the CMB is so uniform, we must also explain why its small fluctuations
are correlated on apparently acausal scales.
50 1000 10 3 1 0 1 3 10 1000
now 1.0
0.8
ne
40
co
conformal time [Gyr]
0.6
ht
lig
scale factor
0.4
30
0.2
20
e
er
0.1
ph
le s
10
bb
Hu
0.01
CMB 0.001
reheating
-10
inflation
-20
-30
-40
causal contact
Fig. 1.2. Inflationary solution to the horizon problem. The comoving Hubble
sphere shrinks during inflation and expands during the conventional Big Bang
evolution (at least until dark energy takes over). Conformal time during inflation
is negative. The spacelike singularity of the standard Big Bang is replaced by
the reheating surface: rather than marking the beginning of time, = 0 now
corresponds to the transition from inflation to the standard Big Bang evolution.
All points in the CMB have overlapping past light cones and therefore originated
from a causally connected region of space.
early times.
In an expanding universe, a shrinking comoving Hubble sphere implies
" #
d 1 1 H H
(aH) = + 1 < 0 <1. (1.4)
dt a H2 H2
We will take the slow evolution of the Hubble parameter, < 1, as our
definition of inflation. This definition includes, but is not limited to, the
dynamics of a slowly rolling scalar field (see 2.2.1). In the de Sitter limit,
0, the space grows exponentially,
a(t) eHt , (1.5)
with H const.
Inflationary expansion requires a somewhat unconventional matter con-
tent. In a spatially-flat FRW universe supported by a perfect fluid, the
Einstein equations lead to the Friedmann equations
2
3Mpl H2 = , (1.6)
2
6Mpl (H + H 2 ) = ( + 3P ) , (1.7)
where and P are the energy density and pressure of the fluid. Combining
(1.6) and (1.7), we find
2
2Mpl H = ( + P ) , (1.8)
and hence
3 P
= 1+ . (1.9)
2
Inflation therefore occurs when P < 13 , corresponding to a violation of
the strong energy condition (SEC).1 One simple energy source that can drive
inflation is a positive potential energy density of a scalar field with negligible
kinetic energy, but we will encounter a range of alternative mechanisms.
With the new cosmology the universe must have been started off in some
very simple way. What, then, becomes of the initial conditions required by
dynamical theory? Plainly there cannot be any, or they must be trivial. We
are left in a situation which would be untenable with the old mechanics.
If the universe were simply the motion which follows from a given scheme
of equations of motion with trivial initial conditions, it could not contain
1
For a perfect fluid, the SEC states that + P 0 and + 3P 0.
20 1 Inflation: Theory and Observations
Inflation not only explains the homogeneity of the universe, but also pro-
vides a mechanism to create the primordial inhomogeneities required for
structure formation [1217]. This process happens automatically when
we treat the inflationary de Sitter phase quantum mechanically. Here,
we briefly sketch the quantum generation of primordial fluctuations. We
also present the modern view of inflation as a symmetry breaking phe-
nomenon [50, 51]. For more details, see Appendices B and C.
Fig. 1.3. Time-dependent background fields m (t) introduce a preferred time slic-
ing of de Sitter space.
The remaining metric fluctuations g00 and g0i are related to by the
Einstein constraint equations. The dynamics of the coupled Goldstone-
metric system can therefore be described by alone.
A second description of the same physics is sometimes convenient, espe-
cially in the cosmological context. First, we note that, for purely adiabatic
fluctuations, we can perform a time reparameterization that removes all
matter fluctuations, m 7 0. This takes us to comoving gauge, where the
field has been eaten by the metric g . The spatial part of the metric
can now be written as
R = H + , (1.14)
where the ellipses denotes terms that are higher order in . This links the
comoving curvature perturbation R with the Goldstone boson of sponta-
neous symmetry breaking during inflation [55, 56].
22 1 Inflation: Theory and Observations
where
y 2 2Mpl
2
. (1.18)
c2s
The field R is therefore massless, implying as we shall see that it is
conserved on superhorizon scales [55].
For simplicity, we will assume that and cs are nearly constant, so that
the overall normalization of the action can be absorbed into the definition
1.2 Primordial Perturbations 23
c2s k 2
vk + 3H v k + vk = 0 . (1.20)
a2
This is the equation of a simple harmonic oscillator with a friction term
provided by the expanding background. The oscillation frequency depends
on the physical momentum and is therefore time-dependent:
cs k
k (t) . (1.21)
a(t)
At early times (small a), k H for all modes of interest. In this limit,
the friction is irrelevant and the modes oscillate. However, the frequency of
each mode drops exponentially during inflation. At late times (large a), the
dynamics is dominated by friction and the mode has a constant amplitude.
We say that the mode freezes at horizon crossing, i.e. when k (t? ) = H or
cs k = aH(t? ). It is these constant superhorizon fluctuations that eventually
become the density fluctuations that we observe in the CMB or in LSS (see
fig. 1.4).2
2
Recall that we are assuming adiabatic initial conditions. The presence of entropy per-
turbations, as in multi-field models, can complicate the relation between the curvature
perturbations at horizon crossing and the late-time observables.
24 1 Inflation: Theory and Observations
comoving
scales
evolution
zero-point
fluctuations
time
horizon reheating horizon CMB today
exit re-entry
Fig. 1.4. The evolution of curvature perturbations during and after inflation: the
comoving horizon (aH)1 shrinks during inflation and grows in the subsequent
FRW evolution. This implies that comoving scales (cs k)1 exit the horizon at
early times and re-enter the horizon at late times. In physical coordinates, the
Hubble radius H 1 is constant and the physical wavelength grows exponentially,
a(t) eHt . For adiabatic fluctuations, the curvature perturbations R do not
evolve outside of the horizon, so the power spectrum PR (k) at horizon exit during
inflation can be related directly to CMB observables at late times.
an ordinary harmonic oscillator. The operators a k play the role of the anni-
hilation operators of the quantum oscillators. The vacuum state is defined
by ak |0i = 0. The oscillation amplitude will experience the same zero-point
vk vk0 |0i = (2)3 |vk |2 (k + k0 ),
fluctuations as an oscillator in flat space, h0|
where
1 1
|vk |2 = . (1.23)
a3 2k
The factor of a3 arises from the physical volume element in the Lagrangian
(1.17)note that the Fourier mode vk was defined using the comoving coor-
dinates rather than the physical coordinates. The second factor, 1/(2k ), is
the standard result for the variance of the amplitude of a harmonic oscillator
in its ground state. (In inflation, this state is the Bunch-Davies vacuum.)
As long as the physical wavelength of the mode is smaller than the Hubble
radius, the ground state will evolve adiabatically. Eq. (1.23) then continues
to hold and the precise time at which we define the initial condition is not
important. Once a given mode gets stretched outside the Hubble radius, the
adiabatic approximation breaks down and the fluctuation amplitude freezes
1.2 Primordial Perturbations 25
at
1 1 1
|vk |2 = , (1.24)
2 a3? cs k/a?
where a? is the value of the scale factor at horizon crossing,
cs k
=H . (1.25)
a?
Combining (1.25) and (1.24), we get
1 H2
|vk |2 = , (1.26)
2 (cs k)3
k3 1 H4
2R (k) 2 PR (k) = . (1.28)
2 8 2 Mpl
2
|H|cs
Since the right-hand side is supposed to be evaluated at horizon crossing,
cs k = aH, any time dependence of H and cs translates into a scale de-
pendence of the power spectrum. Scale-invariant fluctuations correspond to
2R (k) = const., and deviations from scale invariance are quantified by the
spectral tilt
d ln 2R
ns 1 = 2 , (1.29)
d ln k
where we have defined two additional expansion parameters,
cs
and . (1.30)
H Hcs
Inflationary backgrounds typically satisfy {, |
|, ||} 1 and hence predict
ns 1. Inflation would not end if the slow-roll parameters vanished, so
importantly we also expect a finite deviation from perfect scale-invariance,
ns 6= 1.
26 1 Inflation: Theory and Observations
e+pH + , (1.40)
Fig. 1.5. CMB anisotropies as observed by the Planck satellite. Red (blue) spots
are hotter (colder) than the average temperature, reflecting density variations at
recombination.
3
Recall that in 1.2.2 we computed a quantum average. This is related to the ensemble
average after decoherence turns the quantum state into a single classical state of the
ensemble: see e.g. [6569].
1.3 Cosmological Observables 29
where the transfer function T` (k) describes both the evolution of the initial
fluctuations from the moment of horizon entry to the time of recombination,
as well as the projection from recombination to today [57, 62]. Since the
transfer function depends only on known physics it is computable using a set
of coupled Einstein-Boltzmann equations for the primordial plasma [70,71].
The knowledge of T` (k) allows us to use the observed C` as a probe of the
initial conditions R (k). The theoretical curve in fig. 1.6 assumes a nearly
scale-invariant spectrum as predicted by inflation.
30 1 Inflation: Theory and Observations
6000
5000
4000
3000
2000
1000
500 200
250 100
0 0
-250 -100
-500 -200
2 5 10 20 500 1000 1500 2000 2500
a2,`m + a2,`m
E(n) aE,`m Y`m (n) , aE,`m , (1.51)
2
a2,`m a2,`m
B(n) aB,`m Y`m (n) , aB,`m . (1.52)
2i
The E-modes are parity-even, while the B-modes are parity-odd. Roughly,
we can think of the E-mode as the gradient of a scalar and the B-mode as
the curl of a vector. Typical E- and B-patterns are shown in fig. 1.7. Given
T , E and B, we can form several types of correlation functions
E-mode B-mode
(grad) (curl)
Fig. 1.7. Examples of E-mode and B-mode patterns of CMB polarization. While
the E-mode patterns are mirror-symmetric, the B-mode patterns are interchanged
under reflection about a line going through the center.
z 1/3
2 2 c
z ) (1 + z) DA (
DV ( z) . (1.56)
H(z)
Comparing CMB and LSS measurements provides important information
about the evolution of the universe after recombination and helps to break
1.4 Current Tests of Inflation 33
non-linear
0.04
0.02
0.00
4 1
In the Planck analysis, As and ns are defined at the pivot scale k? = 0.05 Mpc .
1.4 Current Tests of Inflation 35
2
b h 0.02207 0.00067 0.02207 0.00054 0.02214 0.00048
2
c h 0.1196 0.0061 0.1198 0.0052 0.1187 0.0034
0.683 0.040 0.685 0.033 0.692 0.021
0.097 0.080 0.091 0.027 0.092 0.026
9
10 As 2.23 0.32 2.20 0.11 2.20 0.11
ns 0.962 0.019 0.959 0.014 0.961 0.011
Table 1.1. Parameters of the CDM baseline model (with 2 errors). The first
four parameters describe the composition of the universe, the last two its initial
conditions. The BAO data improves the constraint on . The small-scale CMB
data hardly affect the constraints but help with a characterization of foregrounds,
which becomes essential when going beyond the CDM model.
0.25
0.20 c
on
ca
con ve
vex
0.15
small-field
0.10 large-field
(chaotic)
0.05
l
natura
0.00
0.94
0 94 0.96
0 96 00.98
98 1.00
Geometry
Inflation very effectively solves the flatness problem [18]. The baseline anal-
ysis of Planck has therefore fixed the curvature parameter to be vanishing,
K = 0. On the other hand, including K in the fit allows a test of
this key prediction of inflation. Table 1.2 shows the constraints on the pa-
rameter K , after marginalizing over the other parameters of the CDM
model. Here, the BAO data plays a crucial role in breaking the geometric
degeneracy between m and H0 and reducing the error on K by an or-
der of magnitude. Even at this new level of precision the observable patch
of the universe is consistent with spatial flatness. Planck has also tested
the isotropy assumption [102]. Except perhaps on the largest scales, the
universe indeed seems to be statistically isotropic.
Table 1.2. Constraints on the geometry of the universe (with 2 errors). The
inclusion of BAO data plays an important role.
Scalar Fluctuations
The observations of the primordial scalar fluctuations are in striking agree-
ment with the predictions of inflation, both qualitatively and quantitatively:
when they re-enter the horizon. All modes with the same wavenumber
k, but possibly distinct wavevectors k, therefore start their evolution
at the same time. This phase coherence allows for constructive in-
terference of the modes and yields acoustic oscillations in the CMB.
Alternative mechanisms for structure formation involving topological
defects (e.g. cosmic strings, see 4.5.2) source perturbations with in-
coherent phases, smearing out the peaks [105], and are therefore ruled
out by the CMB observations. Isocurvature fluctuations also destroy
some of the phase coherence5 and are hence significantly constrained
by the data (see below).
0.4
0.2
0.0
-0.2
-0.4
. Power law spectrum.We have seen above that slow-roll inflation pre-
dicts a power law spectrum with a percent-level deviation from perfect
scale-invariance, which Planck has detected at high significance. At
second order in the slow-roll expansion, inflation predicts a small cor-
rection to the power law spectrum
ns 1+ 1 s ln(k/k? )
2 k 2
R (k) = As . (1.60)
k?
The data is not yet precise enough to detect the expected running of
the spectrum, s (ns 1)2 , and a detection of running at a level
5
In contrived scenarios, causal evolution inside the horizon yields isocurvature perturba-
tions that lead to acoustic peaks [106] see the review [107].
38 1 Inflation: Theory and Observations
Table 1.3. Constraints on tensor modes and on deviations from the power law
spectrum (with 2 errors).
Tensor Fluctuations
Tensor modes contribute to the CMB temperature power spectrum in a
specific way and are therefore constrained by the Planck analysis. Fig. 1.10
shows the current constraints on the parameters ns and r. Marginalizing
over ns gives an upper limit on the tensor-to-scalar ratio [9]
This constraint is at the limit of what can be achieved with CMB tempera-
ture data alone [108]. To probe smaller values of r requires measurements of
CMB polarization: as we explained in 1.3.2, B-modes are a unique signa-
ture of inflationary tensor modes. The BICEP2 collaboration has recently
reported a detection of primordial B-modes. We discuss this result in 1.4.3.
Non-Gaussianity
The CMB power spectrum in fig. 1.6 reduces the Planck data from about 50
million pixels to 103 multipole moments. This enormous data compression is
1.4 Current Tests of Inflation 39
where |0i denotes the vacuum state and R is the quantum operator associ-
ated with the field R. In principle, there is more information in the vacuum
expectation values of higher-order n-point functions. Schematically, we can
write these as the following path integral
Z
h| Rk1 Rkn |i [DR] Rk1 Rkn eiS[R] ,
(1.63)
6
This theorem can be interpreted as a Ward identity associated with the non-linearly
realized dilatation symmetry of the background [56, 118120], and is the analog of the
Adler zero in pion physics.
1.4 Current Tests of Inflation 41
squeezed
equilateral
1.0
folded
1.0
Fig. 1.12. Bispectrum of the local ansatz. The signal is peaked for squeezed
triangles.
squeezed
equilateral
1.0
1.0
1.0
ically, with greater than 70% correlation with the orthogonal template
for 3.1 . A . 4.2.
The Planck collaboration has reported the following constraints on the am-
plitudes of the templates (1.66), (1.67) and (1.68) [10]:
local
fNL = 2.7 5.8 , (1.72)
equil
fNL = 42 75 , (1.73)
ortho
fNL = 25 39 . (1.74)
Non-Adiabaticity
As we have seen, single-field inflation predicts initial fluctuations that are
adiabatic. Adiabatic perturbations have the property that the local state of
matter (determined, for example, by the energy density ) at some spacetime
point (t, x) of the perturbed universe is the same as in the background
universe at some slightly different time t+(t, x). That is, some parts of the
universe are ahead and others behind in the evolution. At recombination,
1.4 Current Tests of Inflation 43
orthogonal
slow-roll
DBI
All matter perturbations therefore have the same density contrast (e.g. b =
c ) and are proportional to the radiation perturbations (e.g. c = 34 ). For
adiabatic initial conditions, all species fluctuate synchronously and lead to
the curvature perturbation R.
In multi-field inflation, it is possible to generate so-called isocurvature
perturbations, where an overdensity in one species compensates for an un-
derdensity in another, resulting in no net curvature perturbation. For exam-
ple, we can define the following isocurvature perturbation for dark matter
and photons,
S c 43 . (1.76)
If this field were significantly different from zero it would lead to a measur-
able effect in the CMB power spectrum.
We digress briefly to describe two classic distinctions between isocurva-
ture perturbations and curvature perturbations, at the level of the acoustic
peaks [125, 126]. The first distinction involves the angular positions of suc-
cessive peaks. Adiabatic perturbations from single-field inflation have fixed
44 1 Inflation: Theory and Observations
amplitude outside the horizon, and begin to evolve upon entering the hori-
zon. The resulting evolution may be thought of as a cosine mode. The
curvature perturbations sourced by cosmic defects, in contrast, have negli-
gible amplitude as they enter the horizon, and grow subsequently through
causal processes. The result is typically a sine mode. These two cases make
different predictions for the angular positions of subsequent peaks, which
are in the ratio 1 : 2 : 3 in the cosine case, and 1 : 3 : 5 in the sine case. The
relative heights of even and odd peaks provide another means of testing adi-
abaticity. Acoustic peaks corresponding to compression waves namely,
the odd peaks are enhanced compared to even peaks in the adiabatic
case, but suppressed compared to even peaks in the isocurvature case.
Definitive evidence against isocurvature models involving causal evolution
inside the horizon, without an inflationary phase, comes from measurements
of CMB polarization. A characteristic signature of these models is that the
temperature and E-mode polarization perturbations are positively corre-
lated on large angular scales [125], while in inflation these perturbations are
anti-correlated. The measurement of TE anti-correlation on superhorizon
scales [127] shows that superhorizon adiabatic perturbations were present
when the CMB decoupled.
Although purely isocurvature perturbations are now ruled out, it is possi-
ble that the observed anisotropies originate from a combination of adiabatic
and isocurvature perturbations. To quantify the isocurvature contribution,
it is conventional to define the relative amplitude of the power spectra of
the isocurvature field and the curvature perturbation
PS
. (1.77)
PR
Assuming that S and R are uncorrelated (motivated by axion isocurvature
models[128130]), Planck has constrained this ratio [9],
0 < 0.036 . (1.78)
The constraint strengthens if S and R are perfectly correlated (as in curva-
ton isocurvature models [131, 132]),
+1 < 0.0025 . (1.79)
Observing an isocurvature contribution to the primordial fluctuations is
another way to rule out single-field inflation, since only the presence of ad-
ditional light fields can give rise to non-adiabaticity. Unfortunately, the
amplitude of the signal depends on the post-inflationary evolution: the
primordial perturbations become adiabatic if the particles produced after
inflation reach a suitable thermal equilibrium [133]. Correspondingly, ob-
servable isocurvature is possible only when one or more particle species has
an abundance determined by physics beyond thermal equilibrium.
1.4 Current Tests of Inflation 45
55
0
60
65
1.8
55
0
60
65
0.3
50 0 50
Right ascension [deg.]
Fig. 1.15. E-mode and B-mode maps measured by the BICEP2 experiment (figure
adapted from [134]). An excess over the lensing B-mode is detected with high
signal-to-noise.
The BICEP experiment was designed specifically to search for the pri-
7
Even if the BICEP2 measurement turns out to be correct in every detail, further ex-
perimental and theoretical work will be required to exclude alternative explanations
for a spectrum of primordial gravitational waves, although no alternative is nearly as
compelling as the inflationary prediction. One way to confirm the inflationary origin of
the signal would be by establishing the superhorizon nature of the B-modes at recom-
bination [135].
46 1 Inflation: Theory and Observations
mordial B-mode signal on degree angular scales. Located at the South Pole,
it observed a small and exceptionally clean patch of the sky, the Southern
Hole. The observation frequencies were chosen to avoid contamination from
synchrotron radiation and from emission by dust. The first version of the
experiment, BICEP1, observed at two frequencies: 100 GHz and 150 GHz.
Collecting data from 2006 to 2008, it obtained the first significant upper
limit on r from polarization measurements alone [136]
r < 0.73 (95% limit) . (1.80)
BICEP2 observed at only one frequency, 150 GHz, but with ten times as
many detectors as BICEP1. Data was taken over three seasons from 2010
to 2012. The final polarization maps are shown in fig. 1.15. Even by eye,
the B-mode pattern is clearly visible! The derived B-mode power spectrum
is shown in fig. 1.16. The best-fit value for the tensor-to-scalar ratio is8
r = 0.2+0.07
0.05 . (1.81)
The null hypothesis r = 0 is rejected at almost 7.
The result in (1.81) was obtained without any foreground subtraction.
The signal is large enough to dominate over available estimates of the po-
larized foregrounds (synchrotron and dust), but a more direct and convinc-
ing exclusion of a foreground explanation would be a detection of B-modes
at a second frequency that confirms the expected thermal spectrum of the
cosmological signal. Using foreground models to correct for any residual
foreground contamination tends to reduce the maximum likelihood value of
r, but not by enough to seriously weaken the significance of the claimed
detection.
The BICEP team took exceptional care to test for systematic errors. They
performed a large suite of so-called jackknife tests. Here, the data is split
according to various criteria and then the difference of the two sets is taken.
The signal will cancel but any systematic effects that vary between the two
sets may remain and can therefore be identified. No failures of any jackknife
tests have been reported.
8
The central value of r claimed by BICEP2 seems somewhat in tension with the Planck
upper bound (1.61). This issue is currently under active investigation [137, 138], so we
will limit ourselves to a few remarks on this issue. Most importantly, the measured
r has not yet stabilized to changes in the analysis. For example, there is still a large
spread in the maximum likelihood values of r for different models of foregrounds, roughly
0.12 < r < 0.21. We therefore caution against a premature judgement of the issue. Even
if the apparent tension survives further scrutiny, it has to be recognized that Planck and
BICEP are sensitive to tensors in different ways. While BICEP measures tensors quite
directly via their imprint on B-mode polarization, Planck constrains the combined effect
of tensors and scalars on the temperature power spectrum. The Planck constraint on
r is therefore model-dependent and weakens if the scalar power is suppressed on large
scales.
1.5 Future Tests of Inflation 47
BICEP2 CBI
BICEP1 Boomerang
QUAD DASI
QUIETQ WMAP
QUIETW CAPMAP
g
sin
len
Multipole
Fig. 1.16. B-mode power spectrum measured by BICEP2, as well as 95% upper
limits from several previous experiments (figure adapted from [134]). Also shown
is the best-fit theoretical curve for r = 0.2. This has two components: one from
primordial tensors that peaks around ` 80, and one from the lensing conversion
of E-modes that peaks around ` 1000.
section, we discuss what one can hope to learn from measuring the pri-
mordial perturbations with increased precision and over a wider range of
scales.
9
The NEC states that the stress tensor satisfies T n n 0, for all null vectors n .
For a perfect fluid, the NEC reduces to + P 0.
1.5 Future Tests of Inflation 49
s = 162 6
+ , (1.84)
where /(H ). Measuring s would test the consistency of the slow-
roll expansion. However, because the running is second order in slow-roll,
we expect it to be small, s (ns 1)2 . Current bounds on s are still
two orders of magnitude larger than this target, but future galaxy surveys
may allow such a measurement [157] (see also [158]). Any detection of a
larger level of running would be a challenge for slow-roll inflation and would
require additional physics to explain.
1.5.3 Non-Gaussianity
The constraints on primordial non-Gaussianity from the CMB have almost
reached their limit. Silk damping of the small-scale anisotropies prohibits
using multipoles larger than `max 2000 to extract information about initial
conditions. This limits the number of modes available in the CMB to
`max 2
CMB
N 106 , (1.85)
`min
which is nearly saturated by the recent Planck measurements.
More modes are in principle accessible through large-scale structure mea-
surements. This is because galaxy surveys probe the three-dimensional
cosmic density field, while the CMB is only a two-dimensional projection.
Hence, while N CMB `2max for the CMB, we have N LSS kmax 3
for LSS
(but see [159, 160]), where kmax is associated with the smallest scale that is
both measurable and under theoretical control. Pushing to smaller scales
(larger kmax ) therefore increases rather dramatically the amount of informa-
tion that can be extracted from the data. The total number of (quasi-)linear
modes in LSS is estimated to be
kmax 3
LSS
Nlinear 109 , (1.86)
kmin
where we have taken kmax 0.1 Mpc1 and kmin 104 Mpc1 . Al-
though this shows the great potential of LSS observations, it assumes that
we measure the entire volume at low redshift. More realistically, we have
kmin 103 Mpc1 (e.g. for the Euclid mission), and hence [161]
kmax 3
Euclid
Nlinear 106 , (1.87)
kmin
which is comparable to the result (1.85) for the CMB. However, while `max
for the CMB cannot be extended, for LSS kmax might be pushed to larger
50 1 Inflation: Theory and Observations
3 3 local
Name zmax V (Gpc/h) ng (Mpc/h) kmax h/Mpc fNL
3
SDSS LRG 0.315 1.48 1.36 10 0.1 5.62
3
BOSS 0.35 5.66 0.27 10 0.1 3.34
3
Big-BOSS 0.5 13.1 0.30 10 0.1 2.27
3
HETDEX 2.7 2.96 0.27 10 0.2 3.65
3
CIP 2.25 6.54 0.50 10 0.2 1.03
3
EUCLID 1.0 102.9 0.16 10 0.1 0.92
3
WFIRST 1.5 107.3 0.94 10 0.1 1.11
Table 1.4. Compilation of current and future LSS surveys. Here, zmax refers to the
maximal redshift of the survey, V is the survey volume and ng is the mean comoving
local
number density of objects. The projected errors on fNL are from measurements
of the galaxy bispectrum. (Data collected by Donghui Jeong.)
This possibility is now highly constrained, and further data from Planck
and LSS surveys (see table 1.4) has the potential to rule out the natural
parameter space of a large class of multi-field models.
As we explain in Appendix B, a similar threshold exists for equilateral
non-Gaussianity [175]:
equil
fNL O(1) . (1.89)
Not seeing a signal at the level of (1.89) would allow us to conclude that the
UV-completion of the effective theory is slow-roll inflation, up to pertur-
bative higher-derivative corrections. Conversely, a detection of equilateral
equil
non-Gaussianity with fNL > O(1) would imply that the theory has to be
UV-completed by something other than slow-roll inflation, such as DBI in-
1.5 Future Tests of Inflation 51
1
More details on effective field theory can be found in [177180].
52
2.1 Principles of Effective Field Theory 53
Inflation?
Gravity
Fig. 2.1. Effective field theories in particle physics. Both Fermi theory and
general relativity are non-renormalizable and are interpreted as effective theories.
m < , are included in the effective theory, while heavy particles , with
masses M > , are integrated out, in a sense that we will make precise.
light heavy
Fig. 2.2. Effective field theories describe the physics of light degrees of freedom
below a cutoff scale . We arrive at these theories either by integrating out the
heavy fields (if the complete UV theory is known) or by parameterizing their effects
(if the UV theory is not known or is not computable). In the latter case, symmetries
inform the choice of allowed interactions.
where Ll (Lh ) describes the part of L involving only the light (heavy) fields,
and Llh includes all interactions involving both sets of fields. The Wilsonian
effective action Seff is defined via a path integral over the heavy modes (and
over the high-frequency contributions of the light fields):
Z
iSeff []
e = [D] eiS[,] . (2.3)
In practice, the effective action is rarely found by performing the path inte-
gral. Instead a so-called matching calculation order by order in perturbation
theory is usually more practical [177180].
In the classical approximation, performing the path integral over the
heavy modes corresponds to using the equations of motion to eliminate the
heavy field . In the language of Feynman diagrams, this is the tree-level
approximation. The complete path integral, however, also includes loops of
the heavy fields. These loops describe how the heavy degrees of freedom
are eliminated from the quantum theory. The result is usually non-local,
meaning that it contains terms such as ( + M 2 )1 . However, at low
energies, E M , these terms can be expanded in derivativesfor
example,
2 1
( + M ) = 1 + 2 + , (2.4)
M2 M
and the EFT becomes approximately local. In other words, the effective
2.1 Principles of Effective Field Theory 55
X Oi []
Leff [] = Ll [] + ci (g) , (2.5)
i M i 4
1 1 1 1 1 1
L[, ] = ()2 m2 2 4 ()2 M 2 2 g2 2 . (2.6)
2 2 4! 2 2 4
2
An important exception is an accidental symmetry: if all operators in the UV theory
violating a symmetry S are irrelevant in the sense of the renormalization group (RG),
then S is an approximate symmetry of the low-energy theory. However, in this case the
smallness of the S-violating terms is not mysterious: it is simply a consequence of RG
flow.
3
Operators of dimension equal to four are called marginal operators. Quantum correc-
tions decide if a marginal operator is relevant or irrelevant in the IR.
4
This is a simplified version of the example studied in [180].
56 2 Inflation in Effective Field Theory
m2R = + + , (2.8)
R = + + , (2.9)
5
Of course, there are similar diagrams with the light particle running in the loop. For
simplicity, we hide those terms in the ellipses of eqs. (2.8) and (2.9).
2.1 Principles of Effective Field Theory 57
can also be computed in a loop expansion; e.g. the coupling of the operator
6 is
c1 = + O(g 3 ) + . (2.13)
The trouble with light scalars.From (2.10) we see that the effective mass
mR gets a large contribution from the mass of the heavy field M . Having
a light scalar field in the EFT is therefore unnatural [182] in the sense that
large quantum corrections, O(M ), must be canceled by a large bare mass m
with opposite sign to achieve mR M (see 2.1.2 for a detailed discussion).
This is a real problem, since for natural values of mR the fields are not
even part of the low-energy EFT! The apparent need to fine-tune the Higgs
boson mass is the famous electroweak hierarchy problem of the Standard
Model. As we will see in 2.3, a qualitatively similar (but quantitatively
less dramatic) hierarchy problem exists in inflationary models driven by
scalar fields. Notice that because the loop correction is proportional to the
coupling g, a hierarchy mR M can be natural if we have reason to believe
that g 1 (and m M ). Thus, the strength of the coupling between the
light and heavy fields is a critical parameter, and symmetry structures in
the UV theory that achieve g 1 play an important role in discussions of
light scalar fields (see 2.1.3).
Decoupling.All the divergences in (2.10) and (2.11) can be absorbed into
a renormalization of the parameters of the Lagrangian. Moreover, in the
limit M the effects of the heavy particles disappear completely. This
decoupling of UV physics [183] ensures that the physical effects of massive
particles are suppressed for large M . Decoupling is one reason that the
Standard Model of particle physics was constructed by focusing only on
renormalizable theories, although today we view it as an effective theory.
dimensionless Wilson coefficients. Eq. (2.14) will be our starting point for
discussing inflation in EFT.
Several comments are important at this stage. First, making assump-
tions about the symmetries of the UV theory can be non-trivial: not all
low-energy symmetries have to admit ultraviolet completions. In 2.1.4, we
discuss this issue in the context of string theory. Second, in writing (2.14)
we have introduced an energy scale and a collection of dimensionless co-
efficients ci .6 It is clearly important to understand how to assign values for
the scale and the coefficients ci . The guiding principle for this undertak-
ing, and more generally for the construction and interpretation of effective
field theories, is naturalness.
2.1.2 Naturalness
Naturalness arguments work in two directions, from the top down and from
the bottom up. We will discuss these two aspects of naturalness in turn.
Top-down naturalness
The top-down version of naturalness asserts that the Wilson coefficients in
an EFT will be of order unity if the cutoff is chosen to be of order the
characteristic mass scale M of the UV theory.7
We emphasize that the coefficients in question are those of operators Oi
allowed by the symmetries of the UV theory: approximate or exact symme-
tries of the ultraviolet theory can lead to small or vanishing coefficients for
the corresponding operators in the effective theory. This top-down version
of naturalness is merely a formalization of the expectation of genericity.
When the ultraviolet theory is computable, naturalness is hardly needed, as
the effective theory can be constructed directly by integrating out the heavy
modes, and the Wilson coefficients can be obtained explicitly in terms of
the parameters of the UV theory. However, this favorable circumstance is
very rare, and in particular it does not arise in presently-studied compact-
ifications of string theory. We often only have partial information about
the UV theory. For instance, we might know the relevant scales, but not
the couplings to all light fields. Top-down naturalness is then widely used
as a systematic framework for guessing how the calculation of the effective
theory would turn out if we were strong enough to perform it.
6 4
At the level of (2.14), the joint rescaling and ci i ci leaves the theory
unchanged. Formalizing this leads to the renormalization group and to the running of
the renormalized couplings with energy.
7
The dimensionless couplings g of the light fields to the heavy fields are assumed to be
of order unity for this purpose. Systematically weaker couplings should be incorporated
by defining a higher effective mass scale, M M/g.
2.1 Principles of Effective Field Theory 59
Bottom-up naturalness
The bottom-up version of naturalness allows one to infer, based on prop-
erties of a given low-energy effective theory, the plausible scale M of new
physics. Here, new physics refers to the scale of the lightest degrees of free-
dom that are part of the ultraviolet theory but not of the effective theory.
Because we generally learn about nature beginning at low energies and pro-
ceeding to higher energies, the bottom-up version of naturalness can be a
predictive tool for the results of experiments.
The logic of bottom-up naturalness arguments is the following: suppose
we are presented with partial information about an EFT like (2.14). Imagine
that we know all the renormalizable couplings of the light fields, but the UV
cutoff and the higher-dimension contributions are unknown this is, for
example, the situation for the Standard Model of particle physics. We would
like to make an educated guess about the size of . In the low-energy theory,
we can calculate loop corrections to the parameters of the renormalizable
Lagrangian as functions of an unknown cutoff scale . The parameters are
said to be bottom-up natural8 as long as their measured values are larger
than the loop corrections. As we extrapolate the EFT to higher energies
and increase , some parameter may become unnatural, and insisting on
natural parameters therefore defines a maximal scale for the effective theory,
= max . Bottom-up naturalness asserts that new physics should appear
at some scale max and modify the effective theory, thereby explaining
the smallness of the measured parameter values. The unnatural alternative
would be that multiple loop and/or bare contributions cancel against each
other for reasons beyond the purview of the effective theory. Bottom-up
naturalness is a formalization of the expectation or more properly, the
hope that this is not the case. In the quintessential example of the Higgs
boson, bottom-up naturalness predicts that new physics should appear at
O(102 103 ) GeV to cut off the quadratic divergence of the Higgs mass mH
and explain why mH = 125 GeV.
The naturalness criterion has been profoundly influential in motivating
physics beyond the Standard Model and it often plays an important role in
inflationary model-building. However, there are reasons to use it with care.
We digress briefly to discuss a few instructive examples illustrating both
successes and failures of the naturalness principle.9
Successes of naturalness.Let us first look at examples where insisting on
natural parameter values has led (or could have led) to the correct physics.
8
We should stress that what we call bottom-up naturalness here is universally referred
to as naturalness, but we find the distinction useful for the present exposition.
9
These comments are based on [184, 185] and on private communications with Nima
Arkani-Hamed.
60 2 Inflation in Effective Field Theory
me = , (2.15)
where re1 . In order for the observed mass of the electron (me
0.511 MeV) to be natural we require < 70 MeV. Indeed, in quantum
field theory, new physics in the form of the positron comes to the
rescue. In [186], Weisskopf showed that virtual positrons surrounding
the electron precisely cancel the linear divergence in (2.15), leaving
only a logarithmic dependence on the cutoff,
me = me ln (/me ) . (2.16)
In the new effective theory, containing both the electron and the
positron, the small electron mass is natural even for large .
. Rho meson.The mass difference between the charged pions and the
neutral pion receives a quantum correction from photon loops
3 2
m2+ m20 = , (2.17)
4
where is the UV cutoff of the effective theory of pions. In order
for (2.17) not to exceed the measured mass splitting, m2+ m20 =
(33.5 MeV)2 , we require < 850 MeV. New physics in the form of
the rho meson with m = 770 MeV comes in at exactly the scale
suggested by naturalness. The charged pion mass is natural in the
new EFT that includes the rho meson.
. Charm quark.Historically, one of the most interesting applications
of the naturalness principle is K 0 -K
0 mixing. In an effective theory
valid below the kaon mass scale, the mass splitting between the KL0
and KS0 states takes the form
mK 0 mK 0 G2F fK
2
L S
= sin2 c 2 , (2.18)
mK 0
L
6 2
in terms of the cutoff , the Cabibbo angle sin c 0.22, and the
kaon decay constant fK = 114 MeV. For (2.18) to be compatible with
the measured splitting (mK 0 mK 0 )/mK 0 = 7 1015 , we require
L S L
< 2 GeV. In fact, new physics in the form of the charm quark, with
2.1 Principles of Effective Field Theory 61
2.1.3 Symmetries
The interplay between symmetry structures in ultraviolet theories and light
scalars in effective theories is crucial for understanding inflation in effective
10
One or two neutrino masses may have the correct scale, but this has not led to a solution
to the cosmological constant problem.
2.1 Principles of Effective Field Theory 63
field theory and string theory, so we now discuss these issues in more depth.
A pivotal question in inflationary model-building in effective field theory
is whether light scalars with m H can be natural (see 2.3). As we
have just explained, whether a given effective theory can be considered
natural depends on the properties of the ultraviolet theory, whether known
or assumed. Symmetries often play a central role in the radiative stability
of the low-energy theory. In this section, we explain this fact for theories
in flat space. In the next section, we will discuss some subtle aspects that
arise in the generalization to ultraviolet completions that include gravity.
SUSY in flat space.We have seen that, in the absence of symmetries,
scalar masses receive loop corrections of the form
m2 2 . (2.22)
There are only a few known ways to protect scalars from these effects. One
elegant possibility is unbroken supersymmetry (SUSY), which obliges boson
and fermion loops to cancel, so that the scalar mass is not renormalized.
However, as we will see, SUSY is necessarily broken during inflation, gener-
ating a mass of the order of the Hubble scale H. Although m H can be
significantly smaller than , it still inhibits successful inflation. Below we
will have more to say about this.
Global symmetries in flat space.Another possibility to stabilize light scalars
is a global internal symmetry. As a concrete example, suppose that the
renormalizable part of the EFT, Ll [], respects the shift symmetry
7 + const. (2.23)
the other hand, it still makes sense to ask whether the fact that the symme-
try is weakly broken in the first place is dictated by some mechanism and
is natural in the top-down sense. In principle, a symmetry can be broken
explicitly by an operator whose coefficient is small purely because of fine-
tuning, and the resulting small parameters are technically natural but not
top-down natural.
Ultraviolet completion.Exact or approximate symmetries of the UV the-
ory can control the sizes of the Wilson coefficients in the non-renormalizable
part of the effective Lagrangian. If the symmetry is weakly broken by the
heavy degrees of freedom, or if the light fields couple only weakly to the
symmetry-breaking terms, then the EFT enjoys an approximate symmetry,
and the Wilson coefficients of all symmetry-breaking operators will be nat-
urally small. This can be seen in our toy model (2.6): the coupling g 2 2
breaks the shift symmetry in the UV. In the EFT, this breaking shows up
through symmetry-breaking irrelevant operators. Since the symmetry is re-
stored in the limit g 0, the Wilson coefficients of all symmetry-breaking
operators in (2.7) must satisfy
2.1.4 Gravity
Gravity plays a fundamental role in any description of cosmology, so our
effective theory must include gravitational degrees of freedom. Moreover,
semi-classical gravity itself has a limited range of validity. At or below the
Planck scale, graviton-graviton scattering violates perturbative unitarity,
and we expect new degrees of freedom to become relevant. In this section,
we discuss to what extent the UV completion of gravity can affect the matter
interactions in the low-energy effective theory.
Gravity as an effective theory.The low-energy degree of freedom of gravity
is the spacetime metric g , whose leading interactions are determined by
2.1 Principles of Effective Field Theory 65
However, while the Yang-Mills action terminates at a finite order, the ex-
pansion of the Einstein-Hilbertaction contains an infinite number of terms,
coming from the expansion of g and g . Gravity is therefore interpreted
as an effective quantum field theory with cutoff = Mpl . The quantum per-
turbation theory of gravitons is organized in terms of the dimensionless ratio
(E/Mpl )2 , where E is the energy of the process, and this theory breaks down
at the Planck scale. At this point either new degrees of freedom become
important (like the massive excitations in string theory; see 3.1) or a non-
perturbative miracle happens (as in asymptotic safety [197]). In the absence
of detailed information about the UV completion, the simplest assumption
is that the low-energy effective action contains all terms that are consistent
with general coordinate invariance. We can organize this as a derivative
expansion,11
" 2
Mpl
Z
Sg = d4 x g M4 + R + c1 R2 + c2 R R
2
1
+ 2 (d1 R3 + ) + , (2.29)
M
where ci and di are dimensionless numbers that may be expected to be
of order unity. For pure gravity, the scale M should also be the Planck
scale Mpl . However, couplings to matter fields might lead to a hierarchy
11 2
There is no R R term, since by the generalized Gauss-Bonnet theorem R
4R R + R R is topological, and equals the Euler characteristic of the space-
time.
66 2 Inflation in Effective Field Theory
between M and Mpl . Note also that the renormalized value of the cosmo-
logical constant M deduced from cosmological experiments is extremely
far from its natural value Mpl . This is, of course, the famous cosmological
constant problem.
In many string theories, the higher-curvature terms in (2.29) can be com-
puted order by order in the 0 and gs expansions detailed in 3.1. An
important example is type IIB string theory in ten dimensions, where one
finds12 [198]
" #
2
10
Z
M10 (3) 1 4
Sg = d X G R+ R + , (2.30)
2 3 25 M 6
where M10 is the ten-dimensional Planck mass, denotes the Riemann zeta
function, with (3) 1.202; R4 is a particular quartic invariant constructed
from the Riemann tensor; and notably the omitted terms include additional
contributions at O(1/M 6 ), and at higher orders, as well as terms that are
subleading in the string coupling gs . The mass M appearing in (2.30) corre-
sponds to the mass of the first excited level of the type II superstring, given
by
4
M2 = 0 , (2.31)
where 0 is the inverse string tension defined in 3.1. This is the proper
physical cutoff scale because the higher-derivative term in (2.30) arises upon
integrating out the massive excitations of the string, which have the mass
spectrum m2 = 4N/0 , N Z. In the regime of weak coupling and weak
curvature where the corrections to the Einstein-Hilbert action are small, the
string scale M is small compared to the Planck scale Mpl , so the higher-
derivative contributions shown above are more significant in string theory
than general reasoning about quantum gravity would suggest. On the other
hand, the coefficient (3)/(325 ) 102 is rather small, illustrating that the
general expectation of order-unity Wilson coefficients should not be viewed
as a precise and immutable law.
Global symmetries in quantum gravity.A number of folk theorems state
that exact continuous global symmetries are impossible in a theory of quan-
tum gravity. Instead, any continuous global symmetry must be merely an
accidental symmetry of the low-energy effective theory, broken by irrelevant
operators at a scale no larger than Mpl . We will briefly recall some of the
arguments; see [199] for a modern discussion of related issues.
12
The ten-dimensional cosmological constant is set to zero in (2.30) because we are dis-
playing the supersymmetric effective action, but the four-dimensional cosmological con-
stant arising upon compactification and supersymmetry breaking is subject to the usual
cosmological constant problem.
2.1 Principles of Effective Field Theory 67
13
Shift symmetries are an important example where the absence of a conserved charge
requires a refinement of the black hole arguments.
14
For example, in 5.4, we will study axions in string theory. An axion with infinite
periodicity (f /Mpl ) enjoys the exact shift symmetry 7 + const., so the
above arguments suggest that such axions are not possible in quantum gravity. And
indeed, direct searches for axions with f & Mpl in parametrically controlled string
compactifications have been unsuccessful [46, 47]. A different argument against axions
with f Mpl appears in [203].
15
Exact Lorentz symmetry is possible in string theory, cf. [204].
68 2 Inflation in Effective Field Theory
findings in string theory, limit the sorts of global symmetries that are allowed
in an ultraviolet completion. Asserting a symmetry structure for the UV
theory and taking natural coefficients for the operators in the resulting
effective action may lead to an effective theory that is consistent at low
energies but cannot be embedded in a theory of quantum gravity. Because
constraints from quantum gravity can play a critical role in determining the
effective action, we view it as prudent to examine any postulated symmetry
structure in a theory of quantum gravity.
Coupling quantum field theory to gravity.Thus far we have discussed flat-
space quantum field theories, as well as purely gravitational theories, but
the theories of interest in cosmology are quantum field theories coupled to
gravity. Let us illustrate this by coupling the toy model of (2.7) to a grav-
itational theory with higher-curvature corrections. The resulting effective
theory takes the form
Here, Oi [g, ] are operators constructed from curvature invariants and from
and its derivatives. In spacetimes where the curvature is small in units of
the cutoff , the only important coupling in Sg, is
Z
(4)
Sg, = d4 x g 2 R , (2.34)
2.1.5 Time-Dependence
To complete our survey of the basic principles of effective field theory, we
need to discuss if and how effective field theory applies to time-dependent
settings such as those arising in cosmology. (For further discussion, see [177,
205, 206].)
2.2 Effective Theories of Inflation 69
where we have allowed for an arbitrary inflaton potential V () (see fig. 2.3).
Fig. 2.3. Example of a slow-roll potential. Inflation occurs in the shaded part of
the potential. In addition to the homogeneous evolution (t), the inflaton experi-
ences spatially-varying quantum fluctuations (t, x).
Inflation ( < 1) therefore occurs when the potential energy of the field
dominates over the kinetic energy, V 21 2 . The kinetic energy stays small
3H||.
and slow-roll persists if the acceleration of the field is small, ||
The conditions for prolonged slow-roll inflation can be expressed as con-
ditions on the shape of the potential [209]:
2 2
V0 |V 00 |
Mpl 2
1 , || Mpl 1. (2.39)
2 V V
2.2 Effective Theories of Inflation 71
16 1
In the WMAP analysis the pivot scale was chosen to be k? = 0.002 Mpc , while for
1
Planck k? = 0.05 Mpc .
72 2 Inflation in Effective Field Theory
You know how sometimes you meet somebody and theyre really nice, so you
invite them over to your house and you keep talking with them and they keep
telling you more and more cool stuff? But then at some point youre like,
maybe we should call it a day, but they just wont leave and they keep talking
and as more stuff comes up it becomes more and more disturbing and youre
like, just stop already? Thats kind of what happened with inflation.
Max Tegmark [210].
V () = 4p p , (2.46)
p2 Mpl 2 Mpl 2
= , = p(p 1) . (2.47)
2
Notice that and do not depend on the scale . Using (2.45), the number
of e-folds occurring in the region ? is found to be
? 2
1
N? , (2.48)
2p Mpl
(a) (b)
(c) (d)
Fig. 2.4. Examples of different classes of slow-roll potentials: (a) chaotic inflation,
(b) natural inflation, (c) hilltop inflation and (d) inflection point inflation. The light
grey regions indication the parts of the potential where slow-roll inflation occurs.
The dark grey regions denote regions of eternal inflation. The figures are not drawn
to scale: (a)+(b) correspond to large-field models ( > Mpl ), while (c)+(d) are
small-field models ( < Mpl ).
For small 0 , hilltop inflation [215] occurs (see fig. 2.4c). The higher-order
terms in (2.58) become important for large values of . They define the
precise value end at which inflation ends, and determine the value of the
cosmological constant in the global vacuum after inflation. We will assume
that end . Mpl . If the higher-order terms in (2.58) are irrelevant when the
pivot scale exits the horizon, then the spectral tilt and the tensor-to-scalar
ratio are [215, 216]
ns 1 = 20 , (2.59)
2 2
end end
r = 2(1 ns )2 eN? (1ns ) 103 . (2.60)
Mpl Mpl
2.2 Effective Theories of Inflation 75
The model has two free parameters: the curvature of the hilltop, 0 , and
the field value at the end of inflation, end .
chaotic
0.1
0.01
l
ura
nat
0.001
hilltop
Again, higher-order terms may become important towards the end of in-
flation, but are assumed to be irrelevant when the pivot scale exits the
horizon. To get enough e-folds of inflation, we require |0 | 1. The special
case 0 = 0 corresponds to inflection point inflation (V000 = 0, see fig. 2.4d):
" #
1 3
V () V0 1 + 0 + 3 + . (2.62)
Mpl 3! 0 Mpl
This type of potential arises in D-brane inflation [217, 218] (see 5.1). The
spectral tilt derived from the potential (2.62) is [218, 219]
r r !
0 0 0 0
ns 1 = 4 cot N? , (2.63)
2 2
r = 1620 . (2.64)
76 2 Inflation in Effective Field Theory
Constraints on the total number of e-folds and the scalar amplitude typically
force 0 to be very small and the tensor signal to be unobservable, r 0.01.
Hybrid inflation.Inflationary models with small-field potentials, such as
(c) and (d) in fig. 2.4, often end through an instability induced by coupling
the inflaton field to an additional waterfall field . The combination of a
slow-roll potential and a waterfall instability is called hybrid inflation [220].
As a simple example, consider the two-field potential (see fig. 2.6)
1
V (, ) = V () + V () + g 2 2 , (2.65)
2
where V () is the slow-roll potential and V () is a potential of symmetry-
breaking type,
1 2 2
V () M 2 . (2.66)
4
We assume that V () M 4 /4, so that the dominant contribution to
the inflationary energy density comes from the false vacuum energy of the
symmetry-breaking potential. The coupling between and induces an
effective mass for the waterfall field that depends on the value of the inflaton,
M2 () = M 2 + g2 . (2.67)
This vanishes at the special point = c M/ g. For > c , the field
is stabilized at = 0, and can be integrated out, so that the theory reduces
to that of single-field slow-roll inflation with Veff () M 4 /4 + V (). As
approaches c from above, becomes light and the effective description
involves both fields. Finally, for < c , the field becomes tachyonic and
ends inflation. Notice that hybrid inflation requires a hierarchy between the
masses of the two fields, V, M 2 . This issue is discussed, and technically
natural examples are constructed, in [221, 222].
2.2 Effective Theories of Inflation 77
2
!
Mpl
Z
4 1 ()2 V ()
p
S= d x
g R , (2.69)
2 2
with potential
4
" r #!2
Mpl 2
V () = 1 exp . (2.70)
4 3 Mpl
17
The Starobinsky model has recently received renewed attentionsee [224] for a super-
conformal generalization and [225, 226] for a no-scale supergravity version.
78 2 Inflation in Effective Field Theory
coupling is the operator 2 R.18 Adding this to the action (2.36), we get
" 2 ! #
2
M
Z
pl 1
S = d4 x g 1 + 2 R ()2 4 , (2.74)
2 Mpl 2 4
where, for concreteness, we have chosen a quartic polynomial for the inflaton
potential. The parameter determines the strength of the non-minimal
coupling to gravity. Again, it is convenient to go to Einstein frame by
performing a conformal rescaling, g = 2 g , with 2 1 + 2 /Mpl 2
.
The action then takes the form
" 2 #
Mpl 1
Z
4 2
p
S = d x g R k()() V () , (2.75)
2 2
where
1 + (6 + 1) 2
k() = , (2.76)
(1 + 2 )2
4
Mpl 4 2
V () = , 2 . 2 (2.77)
4 2 (1 + 2 )2 Mpl
Rp
The canonically-normalized field, = k() d, is
s
6 + 1 p
sinh1 6 + 1 6 sinh1 6 q
p
= .
Mpl 2
1+
(2.78)
For 1, this can be approximated as
r
3
ln(1 + 2 ) , (2.79)
Mpl 2
and the potential becomes
4
" r #!2
Mpl 2
V () = 1 exp . (2.80)
4 2 3 Mpl
18
This interaction played a fundamental role in the revival of Higgs inflation [236].
2.2 Effective Theories of Inflation 79
and the predictions for ns and r are those of (2.73). The predictions for
general were derived in [234, 235]:
32 1 2
ns 1 = , (2.82)
16N? 1 N?
12 6 + 1 1 12
r = + 2 + 2 . (2.83)
N? 6 N?
These results are illustrated in fig. 2.5. We see that the model interpolates
between 4 chaotic inflation (for = 0) and the Starobinsky model (for
1).
where P (X, ) is (so far) an arbitrary function of the inflaton field and
of its kinetic energy X 21 ()2 . The stress-energy tensor arising from
(2.84) corresponds to a perfect fluid with pressure P and energy density
= 2XP,X P , where P,X denotes a derivative with respect to X. The
Friedmann equation and the Klein-Gordon equation are
2 d 3
3Mpl H 2 = 2P,X X P and a P,X = a3 P, , (2.85)
dt
so the inflationary parameter (1.4) becomes
H 3XP,X
= 2 = . (2.86)
H 2XP,X P
dP P,X
c2s = = . (2.87)
d P,X + 2XP,XX
80 2 Inflation in Effective Field Theory
ns 1 = 2 , (2.88)
r = 16cs , (2.89)
where and were defined in (1.30). We saw in 1.4.2 that a small sound
speed leads to observable equilateral-type non-Gaussianity. We will discuss
this further in 2.3.4, where we also emphasize the need to UV-complete
theories such as (2.84). In 5.3, we present DBI inflation [38,40] as a specific
example in string theory.
theory of inflation (2.90) that some irrelevant operators play a crucial role at
low energies, not just for precision observables, but even for the zeroth-order
dynamics. Slow-roll inflation is sensitive even to Planck-suppressed opera-
tors. The next section is devoted to a careful discussion of this important
fact.
19
Variations of these problems arise in most non-slow-roll models as well. For example,
when non-trivial kinetic terms make a rapidly-varying potential innocuous (cf. 2.2.3),
one must still ensure that the necessary kinetic terms are not affected by Planck-
suppressed contributions. The general problem is to arrange that the action, not just
the potential, changes slowly during inflation.
82 2 Inflation in Effective Field Theory
only occasionally been successful. The symmetry options are the same as
discussed in 2.1.3: supersymmetry and/or global internal symmetries. We
will discuss these in turn, but it is worth stating the upshot in advance:
supersymmetry ameliorates but cannot completely solve the problem, while
global symmetry arguments require precise control of Planck-suppressed op-
erators breaking the symmetry, motivating a treatment in quantum gravity.
Supersymmetry.Even if the inflaton is part of a supersymmetric action,
the inflationary background solution spontaneously breaks SUSY, because
the energy density is necessarily positive. Nevertheless, SUSY still limits
the size of radiative corrections, because sufficiently high frequency modes
are insensitive to the effects of the spacetime curvature during inflation.
The cancellation between boson and fermion loops therefore still applies in
the high-energy regime, just as in flat space. On the other hand, modes
with frequencies below the Hubble scale, . H, do experience non-trivial
effects from the expanding background. Boson and fermion propagators
are then modified by the coupling to the spacetime curvature, with mass
splittings within supermultiplets that are typically of order H, and the
corresponding loops no longer cancel. Radiative corrections to the inflaton
mass are therefore naturally of order of the Hubble scale,
m2 H 2 . (2.93)
This is smaller than the correction in (2.91), but not small enough to evade
the eta problem:
1 . (2.94)
This qualitative argument is confirmed in detail by investigations of inflation
in supergravity [238] and in string theory (see 4.2). Hence, although SUSY
ameliorates the eta problem, it does not solve it: successful inflation still
requires fine-tuning of the mass term [239, 240], although much less than in
an EFT without SUSY.
The degree of fine-tuning implied by (2.94) depends to some extent on
the underlying model. In small-field models with , the value of
at horizon crossing is related to the scalar spectral index, 21 (ns 1).
For the Planck best-fit, ns 0.96, this implies 0.02, so the required
fine-tuning is at the percent-level.
Global symmetries.We have discussed global symmetries extensively in
2.1.3. As we explained there, a small scalar mass is bottom-up natural
if the renormalizable part of the Lagrangian (the IR theory) respects an
approximate shift symmetry
7 + const. (2.95)
In other words, the theory contains no relevant or marginal operators that
violate (2.95). Then, loops of the light fields do not drive the scalar mass
2.3 Ultraviolet Sensitivity 83
2
O6 = cVl () , (2.96)
2
where c is a constant, and Vl () consists of the renormalizable terms in the
potential, cf. (2.90). Even if Vl () respects an approximate shift symmetry,
this is broken by O6 . Provided that the inflaton vev is smaller than the
cutoff, < , the operator O6 makes only a small correction to the infla-
tionary potential, V V (). Nevertheless, its effect on the inflaton mass
is significant:
Mpl 2
2c . (2.97)
For c O(1) and < Mpl , the theory again suffers from the eta problem.
(Notice that the overall scale of the potential cancels in (2.97).) Even if the
operator in (2.96) is Planck-suppressed, Mpl , it cannot be ignored in
discussions of the inflationary dynamics.
In a theory with a single real scalar , it is difficult to give a convincing
argument for the absence of couplings of the form (2.96). Note in particular
that if one forbids 2 via a global symmetry under which transforms
linearly, one would simultaneously exclude the kinetic term 21 ()2 . An
influential approach is therefore to take to transform nonlinearly under a
global symmetry (e.g. by taking to be an axion see 5.4), and/or to be
the phase of a complex scalar (cf. e.g. [241]).
Although operators of the simple form (2.96) arise in many ultraviolet
completions (see 4.2 and Chapter 5), more general non-renormalizable in-
teractions are also common, and can give comparable (or larger) effects.
84 2 Inflation in Effective Field Theory
20
Notice that for 4 < < 6, 6= 5, the correction increases for small /. Irrelevant
operators with non-integer dimensions < 6 can therefore dominate the dynamics in
small-field inflation. For an example, see 5.1.
2.3 Ultraviolet Sensitivity 85
Fig. 2.7. Evolution of the inflaton field from the time when modes that are ob-
servable in the CMB exited the horizon, ? , to the end of inflation, end . The total
field displacement is related to the tensor-to-scalar ratio r by (2.105).
Integrating (2.100) from the time N? when modes that are observable in the
CMB exited the horizon, until the end of inflation at Nend 0 (see fig. 2.7),
we get [242] r
Z N?
r(N )
= dN . (2.101)
Mpl 0 8
To evaluate the integral in (2.101), it is useful to define
Z N? s
r(N )
Neff dN , (2.102)
0 r?
[244]), while more typically Neff & 50. Taking Neff & 30, we conclude that21
r 1/2
& . (2.105)
Mpl 0.01
To arrive at a maximally conservative bound in single-field slow-roll infla-
tion, one can assume that slow-roll is valid only while the observed multi-
poles of the CMB exit the horizon, corresponding to Neff 7. This leads
to (cf. [242], which used a smaller Neff because fewer multipoles had been
observed in 1996)
r 1/2
& 0.25 . (2.106)
Mpl 0.01
It is quite a remarkable coincidence that the level of tensors that is experi-
mentally accessible (r & 0.01) is tied to the fundamental scale of quantum
gravity, Mpl .
We strongly caution against viewing = Mpl as an absolute dividing
line: the theoretical challenges of models with > Mpl are shared by
models with slightly smaller displacements. In particular, although gravity
itself becomes strongly coupled around the scale Mpl , parametrically con-
trolled ultraviolet completions of gravity generally involve additional scales
< Mpl . For instance, the string scale and the Kaluza-Klein scale (see 4.1)
are typically well below the Planck scale. Field excursions that are large
compared to those scales raise concerns similar to the super-Planckian issues
we describe below.
Finally, let us emphasize that the Lyth bound (2.105) is a purely kine-
matic statement, relating r to the distance in field space over which the
inflaton moves. Although the bound has profound consequences in the con-
text of effective field theory reasoning about natural Planck-suppressed in-
teractions (see below), the derivation of (2.105) relied in no way on notions
of naturalness, or on a Taylor expansion of the potential.
21
One should not assume that simple models will approximately saturate (2.105): for
example, chaotic inflation scenarios involve displacements roughly four times larger
than required by the bound.
2.3 Ultraviolet Sensitivity 87
scalar inflaton. From this perspective, there are two issues that appear
dangerous at first glance, but are in fact not at all problematic [245].
First, one might worry that super-Planckian displacements of the inflaton
will lead to super-Planckian energy densities, and correspondingly large
gravitational backreaction. This concern is misplaced: the normalization
4
(2.42) of the scalar fluctuations requires that V Mpl . For instance,
2 2
in m chaotic inflation, (2.42) implies that the inflaton mass is small,
m 105 Mpl , so that the energy density never becomes significant even for
super-Planckian fields.
A second concern is radiative stability: do quantum corrections, from
graviton loops and/or loops, destabilize the classical potential V ()? No:
the small value of the inflaton mass22 m is technically natural, because the
theory enjoys a shift symmetry in the limit m 0. Quantum corrections
therefore do not destabilize the potential. In particular, the one-loop cor-
rection from graviton loops is [246]
V V 00 V
= c1 2 + c2 4 , (2.107)
V Mpl Mpl
4
where c1 and c2 are order-one numbers. Because m Mpl and V Mpl ,
this is a small correction.
To summarize, in the low-energy theory of the inflaton and the graviton,
potentials supporting large-field inflation can be radiatively stable, and in
particular free of significant corrections from inflaton-graviton interactions.
Thus, from the bottom-up perspective, large-field inflation is not problem-
atic.
The essence of the problem of large-field inflation is that gravity requires
an ultraviolet completion, and couplings of the inflaton to the degrees of
freedom that provide this ultraviolet completion do not necessarily respect
the symmetry structures needed to protect the inflaton potential in the low-
energy theory. The effects of classical gravity (i.e. backreaction) and of semi-
classical gravity (i.e. graviton loops) are not problematic, but full quantum
gravity effects corresponding to integrating out fields with Planck-scale
or string-scale masses are subtle, and have the potential to be ruinous.23
22
A parallel argument applies to chaotic inflation with a non-quadratic monomial poten-
tial.
23
Some authors have argued that quantum gravity effects are necessarily small when
all energy densities are sub-Planckian. As a general statement about an arbitrary
quantum gravity theory, this is false: the fact that the eta problem appears in string
theory, cf. 4.2, is one simple counterexample, and the diverse failure modes of large-
field models in string theory discussed in 5.4 provide many more. Ignoring quantum
gravity effects purely because all energy densities are small in Planck units amounts to
attributing to the quantum gravity theory underlying our universe a property that is
88 2 Inflation in Effective Field Theory
(a)
(b)
Fig. 2.8. (a): Observable tensor modes require a smooth inflaton potential over
a super-Planckian range. (b): In the absence of symmetries, effective field theory
predicts that generic potentials have structure on sub-Planckian scales, < Mpl .
not seen in string theory. This is a logically consistent position but is very far from
being agnostic about quantum gravity.
2.3 Ultraviolet Sensitivity 89
7 + const. (2.109)
24
Note that to realize an approximate shift symmetry in the low-energy theory, it would
suffice for the inflaton to have weak couplings g 1 to all the degrees of freedom of
the UV completion: the Wilson coefficients in (2.108) would then be suppressed by
powers of g. Equivalently, the effective cutoff scale would become Mpl /g Mpl . The
coupling of the inflaton to any additional degrees of freedom would be weaker than
gravitational [247].
90 2 Inflation in Effective Field Theory
25
The Lyth bound is sometimes misunderstood as the statement that detectable grav-
itational waves imply > 1.0 Mpl , or equivalently upon imposing a legalistic
definition of small-field inflation as inflation with < 1.0 Mpl that detectable
gravitational waves are impossible in small-field inflation. Neither statement is true,
so exhibiting counterexamples has limited utility.
2.3 Ultraviolet Sensitivity 91
26
Specifically, we require that the usual slow-roll parameters are small, and moreover the
parameter defined in (C.94) obeys 1.
92 2 Inflation in Effective Field Theory
2.3.4 Non-Gaussianity
Single-field slow-roll inflation has an approximate shift symmetry27 (2.95)
that constrains inflaton self-interactions in the potential and prevents large
non-Gaussianity: fNL O(, ) 1 [116]. To generate observable levels of
non-Gaussianity requires either higher-derivative interactions or couplings
to extra fields. Both options can be ultraviolet sensitive.
Non-Gaussianity from higher derivatives.When higher-derivative interac-
tions are important, the dynamics deviates significantly from slow-roll. In
2.2.3, we presented P (X) theories as a specific example. We mentioned
that fluctuations propagate with a nontrivial speed of sound,
P,X
c2s = . (2.112)
P,X + 2XP,XX
1 X2
P =X+ + , (2.113)
2 4
27
This symmetry does not have to be fundamental, but may be the result of fine-tuning.
Its presence is motivated by the observed scale-invariance of the primordial fluctuations.
2.3 Ultraviolet Sensitivity 93
fNL 1 1 X + 1 .
equil
(2.114)
c2s 4
equil
On the other hand, |fNL | > 1 can only arise for X & 4 , in which case it is
inconsistent to truncate the expansion in (2.113). Instead, an infinite num-
ber of higher-derivative termsthose proportional to ei in (2.108)become
relevant. Observably-large non-Gaussianity in single-field inflation is there-
fore UV sensitive.28 Special symmetries, such as the higher-dimensional
boost symmetry of DBI inflation [38] (see 5.3), are required to make sense
of the UV completion of (2.113).29
Non-Gaussianity from hidden sectors.As we will see in Chapters 4 and 5,
ultraviolet completions of inflation invariably involve extra fields coupled to
the inflaton. We will collectively denote these fields by . If these fields are
sufficiently heavy (m H), they can be integrated out and only affect the
couplings of the single-field EFT. Light hidden sector fields (m < H), on
the other hand, can affect the inflationary fluctuations and may therefore
leave imprints in cosmological observables.
Although the approximate shift symmetry (2.95) sharply limits the non-
Gaussianity that can arise from self-interactions of the inflaton, the cou-
plings of hidden sector fields are much less constrained, and hidden-sector
self interactions can lead to visible non-Gaussianity, as we now explain. Sup-
pose that the shift symmetry of the inflaton is preserved by the coupling to
a hidden sector field . Then the leading interaction between the hidden
sector and the visible sector is the dimension-five operator [267]
X
O5 =
. (2.115)
This coupling converts any non-Gaussianity in the hidden sector into ob-
servable non-Gaussianity in the inflaton sector.
Under rather natural circumstances, the fluctuations in the hidden sector
can be highly non-Gaussian. For example, suppose that supersymmetry is
28
This issue is also visible in the effective theory of fluctuations [51] (see Appendix B). In
the limit of observable non-Gaussianity, the theory of the fluctuations becomes strongly
coupled below the symmetry-breaking scale [265], and must therefore be UV-completed
below . This is in contrast to the slow-roll limit, where questions about the UV
completion are deferred to scales above .
29
Another class of ghost-free, radiatively stable higher-derivative models is Galileon infla-
tion [266]. In these models the renormalization of the action is protected by the Galilean
symmetry 7 + b x + c, which is a combination of the shift symmetry (2.95) and
a spacetime translation. No candidate for an ultraviolet completion of a Galileon model
in string theory has been proposed, and whether one exists is an open question.
94 2 Inflation in Effective Field Theory
30
Order-one non-Gaussianity in the observed curvature perturbations would correspond
to fNL R 1, not to fNL 1.
3
Elements of String Theory
String theory is the subject of a vast literature.1 Our aim in this section is to
assemble the results that are most relevant for the study of string inflation
(the subject of Chapters 4 and 5), making no pretense of completeness. We
will particularly focus on the four-dimensional effective actions arising in
cosmologically realistic solutions of string theory. Careful attention is paid
to the problem of moduli stabilization, and de Sitter solutions are critically
analyzed.
3.1 Fundamentals
3.1.1 From Worldsheet to Spacetime
An elementary starting point for string theory is the worldsheet action for
a string, which defines a (1+1)-dimensional quantum field theory. We will
begin by describing bosonic string theory, and then turn to string theories
whose worldsheet theories include fermionic fields.
Bosonic string theory.The Polyakov action for a bosonic string propagat-
ing in D-dimensional Minkowski space [284, 285] is
Z
1
SP = 0 d2 hhab a X M ()b X N ()M N , (3.1)
4
where X M , with M = 0, , D 1, are the coordinates in the target
spacetime; a , with a = 0, 1, are the coordinates on the string worldsheet;
hab is an independent metric on the worldsheet; and 20 is the inverse
of the string tension. The action (3.1) describes a two-dimensional field
theory with D scalar fields. At the classical level, this theory is invari-
ant under two-dimensional diffeomorphisms and under the Weyl symmetry
1
The fundamentals of the theory can be found in the classic textbooks [270273], as well
as the lecture notes [274280]. More recent advances are described in [281283].
95
96 3 Elements of String Theory
where R(h) is the Ricci scalar constructed from hab . Expanding the back-
ground fields around a given point, X M = X(0) M
+ X M , one finds inter-
action terms such as hab P GM N (X(0) )X P a X M b X N . The nonlinear
-model defined by (3.3) therefore describes an interacting quantum field
theory. When the gradients of the background fields are small in units of 0
and in particular, when all curvatures are small in string units these
interactions can be treated perturbatively. The corresponding expansion is
known as the -model expansion or the 0 expansion. Absence of anomalies
in the quantum field theory defined by (3.3) requires that the background
fields in the target spacetime obey certain differential equations that can be
obtained order by order in the 0 expansion. Consistency of string theory
at the quantum level on the worldsheet therefore imposes equations of mo-
tion in the target spacetime [287]. Remarkably, the equation of motion for
GM N (X) at leading order in 0 is the Einstein equation!
The equations of motion for the background fields can also be shown to
follow from a D-dimensional spacetime action that parameterizes the in-
teractions of the massless excitations of the bosonic string. The idea is to
construct an effective action in the sense described in Chapter 2: one imag-
ines performing the path integral by first integrating out massive excitations
2
This assumes that there is no topological obstruction to the existence of a metric that
is flat everywhere.
3.1 Fundamentals 97
of the string, leaving an effective action for the massless modes. The theory
that emerges at energies below the string scale Ms takes the form (see [272]
for details)
!
Z
1 1 2(D 26)
SB = 2 dD X G e2 R + 4()2 |H3 |2 + O(0 ) ,
2D 2 30
(3.4)
where D is a coupling constant, and H3 = dB2 is the field strength of the
antisymmetric tensor BM N , or equivalently of the two-form B2 . Although
the effective action (3.4) lacks the good ultraviolet behavior of the full string
theory (it violates perturbative unitarity at E Ms ), it is nevertheless a
convenient way to organize the interactions at energies below the cutoff,
E Ms . The omitted terms of higher order in 0 correspond to higher-
dimension operators, including invariants constructed from the Riemann
curvature of the target space.
In practice, the effective action (3.4) is obtained by computing scattering
amplitudes for strings via a path integral over worldsheets connecting initial
and final states. The path integral involves a sum over surfaces connecting
the initial and final configurations, and the genus of the surface is a loop
counting parameter: worldsheets of Euler number appear in the path
integral with weight
string theories are far more promising, and differ in important details. Most
fundamentally, the worldsheet actions involve additional fermionic terms: in
the simplest case, known as N = (1, 1) worldsheet supersymmetry, the total
action in conformal gauge takes the form [270]
Z
1 2
a M M a
S = SP + SF = d X X
a M i a M .
(3.6)
40
Here, a are two-dimensional Dirac matrices obeying the Dirac (or Clifford)
algebra
{a , b } = 2 ab , (3.7)
and M is a Dirac spinor on the worldsheet that transforms as a vector
under Lorentz transformations in the target space (which correspond to
global symmetry transformations of the worldsheet theory). In terms of the
two independent components of M ,
M
M
M , (3.8)
+
the fermion action takes the form
Z
i 2
M N M N
SF = d
+ + + + M N ,
(3.9)
20
where 12 ( ), with 0 and 1 . The worldsheet fermions
therefore separate into left-moving and right-moving modes. The fermions
M
contribute to the central charge of the worldsheet field theory, so that
the theory defined by S = SP + SF , with SP given in (3.2), has the critical
dimension D = 10.
The action (3.9) for the worldsheet fermions does not completely de-
termine the spacetime spectrum of the theory: one must also specify the
periodicity of the fermions under transport around the closed string world-
M M
sheet. Periodic fermions obeying ( +) = + () are said to be in the
M M
Ramond sector, while antiperiodic fermions with ( + ) = () are
said to be in the Neveu-Schwarz sector. This choice can be made separately
for the left-moving and right-moving fermions, so that there are four pos-
sible sectors: NS-NS, R-R, R-NS, and NS-R. The ten-dimensional effective
actions describing the interactions of massless states of the superstring are
supergravity theories involving additional fermionic and bosonic fields in
comparison to (3.4). Bosonic fields in the target spacetime arise from string
states in the NS-NS and R-R sectors, while the R-NS and NS-R sectors give
rise to spacetime fermions.
To construct a consistent closed string theory with spacetime fermions,
it turns out to be necessary to impose a particular projection, the GSO
3.1 Fundamentals 99
projection, on the spectrum. This entails one further choice: one can per-
form identical GSO projections in the R-NS and NS-R sectors, or opposite
projections. The former choice leads to type IIB string theory, which has
a chiral spectrum in spacetimein particular, the two gravitinos have the
same chirality. The latter choice produces type IIA string theory, which has
a non-chiral spectrum.
Three other consistent superstring theories are known. To arrive at type I
string theory, we consider the worldsheet parity operation , which reverses
the orientation of the string worldsheet, and hence relates left-moving and
right-moving modes. In type IIB string theory, the R-NS and NS-R sectors
have the same spectra, so that worldsheet parity is a symmetry of the theory,
and it is consistent to project the spectrum onto states with = +1. This
operation, which corresponds to gauging the discrete symmetry of world-
sheet parity, leads to a theory of unoriented strings, because for any given
string its orientation-reversed image under is also retained. The projec-
tion removes one of the two gravitinos from the spectrum, yielding a theory
with N = 1 supersymmetry in ten dimensions, the type I string.
The two remaining theories also have ten-dimensional N = 1 supersym-
metry, but have a different structure on the worldsheet. While above we
have discussed theories with left-moving and right-moving bosons, and left-
moving and right-moving fermions, it is also consistent to take the left-
moving sector to be that of the bosonic string, and the right-moving sector
to be that of the superstring. Two supersymmetric heterotic string theories
arise from this construction: the SO(32) heterotic string, and the E8 E8
heterotic string.
The five superstring theories described above are interrelated by a number
of dualities (see fig. 3.1), and correspond to different limits of an underlying
theory which is sometimes called M-theory.
Supergravity limit.The low-energy limit of each of the consistent super-
string theories is a ten-dimensional supergravity theory. We will now de-
scribe the corresponding effective actions for type IIA and type IIB string
theory, focusing on the bosonic fields, which are directly relevant for obtain-
ing classical solutions. The actions for type I string theory and the heterotic
string theories may be found in e.g. [273].
The NS-NS sector of type II supergravity in ten dimensions contains the
metric GM N , the dilaton and the two-form B2 . The action for these fields
is
10
Z
1 2 2 1 2
SNS = 2 d X Ge R + 4() |H3 | , (3.10)
2 2
M-theory
heterotic
IIA
T-dual T-dual
heterotic
IIB
Fig. 3.1. Dualities relating the supersymmetric string theories and M-theory. S-
duality exchanges strong coupling and weak coupling, while in compactification on
a circle, T-duality exchanges momentum and winding.
22 = (2)7 (0 )4 . (3.11)
where
Z
(IIA) 1
SR = d X G |F2 | + |F4 |2
10 2
, (3.13)
42 Z
(IIA) 1
SCS = B2 F4 F4 , (3.14)
42
with Fp = dCp1 and F4 = F4 + C1 H3 . The R-R fields in type IIB
supergravity are a zero-form (scalar) C0 , a two-form C2 , and a four-form C4
with self-dual field strength. The complete action is
(IIB) (IIB)
SIIB = SNS + SR + SCS , (3.15)
where
Z
(IIB) 1 10 2 2 1 2
SR = d X G |F1 | + |F3 | + |F5 | , (3.16)
42 Z 2
(IIB) 1
SCS = C4 H3 F3 , (3.17)
42
3.1 Fundamentals 101
3.1.2 D-branes
In addition to fundamental strings, string theory contains solitonic objects.
Most famous are D-branes, which are charged under the gauge symmetry
of the R-R fields. A Dp-brane is an object with p spatial dimensions that
is charged under Cp+1 via the electric coupling3
Z
SCS = p Cp+1 , (3.22)
p+1
3
Type IIA string theory contains stable Dp-branes with p even, while type IIB string
theory contains stable Dp-branes with p odd. Type I string theory has stable Dp-branes
with p = 1, 5, 9. Dp-branes and D(6 p)-branes are charged under R-R potentials Cp+1
and C7p , whose field strengths Fp+2 and F8p are dual to each other, ?F8p = Fp+2 .
Thus, D(6p)-branes carry magnetic charge under Cp+1 . In string theories with an NS-
NS two-form B2 , fundamental strings carry charge under B2 and are stable. Moreover,
there is an additional soliton, the NS5-brane, which is magnetically charged under B2 .
102 3 Elements of String Theory
where p+1 is the Dp-brane worldvolume and p is the brane charge. The
Chern-Simons action (3.22) is simply a higher-dimensional generalization
R
R the coupling of a charged point particle to a gauge potential, dx A
of
A1 .
A defining characteristic of D-branes is that they are surfaces on which
strings can end. The D in D-brane stands for Dirichlet, referring to the fact
that open strings ending on a D-brane have Dirichlet boundary conditions in
the directions transverse to the brane, i.e. the open string endpoints cannot
leave the D-brane. Open strings have Neumann boundary conditions in the
directions along the spatial extent of a Dp-brane with p > 0: the endpoints
are free to slide along the D-brane.
Quantization of the open strings residing on a D-brane yields a spectrum
of bosonic and fermionic fields living on the worldvolume. At the mass-
less level, one finds scalar fields parameterizing fluctuations of the D-brane
position, a worldvolume gauge field Aa with field strength Fab , and their
superpartners. The effective action for these fields is an important object,
because it encapsulates the low-energy dynamics of the D-brane. Just as the
low-energy effective action for the massless modes of a closed string could be
determined by computing closed string scattering amplitudes in perturba-
tion theory, the low-energy effective action that governs the massless fields
on a D-brane can be derived by computing scattering amplitudes involving
open strings ending on the D-brane. Moreover, by computing amplitudes
in which open strings on the D-brane interact with closed strings, one can
determine the couplings of the D-brane to a closed string background.
A general background solution of type II or type I 4 string theory will in-
volve profiles for all the massless bosonic fields. We would like to understand
the effective action for the light fields on a Dp-brane in such a background.
For simplicity, we will restrict attention to the bosonic sector.
D-brane action.An uncharged p-dimensional membrane moving in a curved
spacetime with metric GM N can be described by the Dirac action, which is
simply a higher-dimensional generalization of the Polyakov action (3.1):
Z
SD = Tp dp+1 det(Gab ) ,
p
(3.23)
where
X M X N
Gab G . (3.24)
a b M N
Here, Gab is the pullback of the metric of the target spacetime, and Tp is
the tension of the membrane.
4
The heterotic string theories contain no R-R fields, and correspondingly lack D-branes.
3.1 Fundamentals 103
leading to the important result that Dp-branes are heavy at weak string
coupling, gs 1. Next, the D-brane tension Tp can be related to the charge
p appearing in (3.22). The stable D-branes of type I and type II string the-
ories are BPS objects, and preserve half of the spacetime supersymmetries.
A BPS Dp-brane (p > 0) corresponds to a higher-dimensional generalization
of an extremal black hole, with tension equal to its charge when expressed
in appropriate units. In our conventions, one finds p = gs Tp .
The Chern-Simons action in the presence of background fields in the tar-
get space and on the D-brane worldvolume takes the form
Z
Cn e F ,
X
SCS = i p (3.29)
p+1 n
where the sum runs over the R-R n-forms of the theory in question, and
only (p+1)-forms contribute to the integral in (3.29). The complete bosonic
5
In a background with constant dilaton , one has gs e = 1.
104 3 Elements of String Theory
3.2 Compactification
For our purposes, a solution of string theory is a configuration of the massless
fields that solves the equations of motion of the effective theory, and hence
3.2 Compactification 105
6 0
The Einstein equations receive corrections in the expansion, which can be important
at large curvatures.
106 3 Elements of String Theory
7
Compactifications that break supersymmetry at the Kaluza-Klein scale (for construc-
tions with stabilized moduli, see e.g. [291293]) present an important alternative to
supersymmetric compactifications.
3.2 Compactification 107
gM N = e2(x) gM N , (3.42)
e2 R
= R 2(D 1)2 (D 2)(D 1)g M N M N . (3.43)
e2
2 = 2 + (D 2)g M N M N . (3.44)
108 3 Elements of String Theory
3.2.2 Moduli
In the simple Kaluza-Klein reduction described above, the breathing mode
corresponding to an overall dilation of the internal space gave rise to a four-
dimensional scalar field u(x) parameterizing spacetime-dependent changes
3.2 Compactification 109
8
A complete treatment can be found in [294], which we follow in this section. See
e.g. [282] for background on Calabi-Yau geometry.
9
Calabi-Yau compactifications of type I string theory, or of the heterotic string, yield
N = 1 supersymmetry. We will primarily discuss type II compactifications.
110 3 Elements of String Theory
O = (1)FL ws , (3.55)
orientifold planes have three or seven spatial dimensions, and are known as
O3-planes and O7-planes, respectively.
Under the action (3.55), the cohomology group H 1,1 can be decomposed
as
H 1,1 = H+1,1 1,1
H , (3.56)
with the subscript denoting the parity of the corresponding two-forms under
the orientifold action. Correspondingly, the basis I , I = 1, . . . , h1,1 for H 1,1
decomposes into a basis for the even eigenspace, i , i = 1, . . . , h1,1+ , and a
1,1
basis for the odd eigenspace , = 1, . . . , h .
To understand the effect of orientifolding on the four-dimensional fields,
we note that v I , I , , C0 are even under the orientifold action, while A ,
bI , cI , B2 (x), C2 (x) are odd. Invariant four-dimensional fields arise from
even ten-dimensional fields expanded in terms of even forms, or from odd
ten-dimensional fields expanded in terms of odd forms. The Kahler form
can be written
J = ti (x)i , (3.57)
1,1
so that the orientifold-invariant K
ahler moduli are the h+ real scalars ti ,
which measure the volumes of two-cycles that are even under the involu-
tion. Similarly, noting that the orientifold action projects out the four-
dimensional two-forms B2 (x) and C2 (x), we have the invariant fields (again
omitting vector contributions)
B2 = b (x) , (3.58)
C2 = c (x) , (3.59)
C4 = i (x)
i . (3.60)
Likewise, the invariant complex structure moduli are , for = 1, . . . , h1,2
.
Finally, and C0 are automatically invariant.
It is important to assemble the invariant scalars into the bosonic compo-
nents of chiral multiplets of four-dimensional N = 1 supersymmetry, i.e. to
determine the proper K ahler coordinates on the moduli space. First of all,
the axion C0 and dilaton combine to form the complex axiodilaton,
= C0 + ie . (3.61)
The complex structure moduli are automatically good Kahler coordi-
nates. The two-form scalars b and c form the complex combination
G c b . (3.62)
To go further, we note that the compactification volume V can be written
in terms of the K
ahler form J as follows:
Z
1 1
V= J J J = cijk ti tj tk , (3.63)
6 X6 6
112 3 Elements of String Theory
where cijk are the triple intersection numbers of X6 . Then, the Kahler
coordinates describing complexified four-cycle volumes are [294]
1 1
Ti cijk tj tk + ii + e ci G (G G)
, (3.64)
2 4
The expression (3.64) is not supposed to be obvious, but we can provide
some intuition by dropping the contribution of G , so that
1
Ti = cijk tj tk + ii . (3.65)
2
Now, we recall that the two-cycle volumes ti are related to the four-cycle
volumes i by
V 1
i = i = cijk tj tk , (3.66)
t 2
so that (3.65) can be recognized as
Ti = i + ii . (3.67)
This is the familiar complexification of four-cycle volumes i by i , i.e. by
the integral of C4 over the corresponding four-cycle. The more involved
expression (3.64) shows that the corresponding proper Kahler coordinate
depends on the vev of the two-form G .10
In summary, the K ahler coordinates on the moduli space are the h1,1
+ com-
1,1
plexified four-cycle volumes Ti (3.64), the h two-form scalars G (3.62),
the axiodilaton (3.61), and the h1,2
complex structure moduli . All
told, a compactification of type IIB string theory on an O3/O7 orientifold
of a Calabi-Yau manifolds leads to h1,1 1,1 1,2
+ + h + h + 1 = h
1,1
+ h1,2
+1
complex moduli scalars in the four-dimensional theory. Further scalar fields
can arise from the open string sector.
3.2.3 Axions
One class of fields deserves special discussion: these are axions, i.e. pseu-
doscalar fields enjoying Peccei-Quinn (PQ) shift symmetries of the form
a 7 a + const. (3.68)
The QCD axion is the original and most famous example of an axion, and
some authors reserve the word axion for this field alone, but we stress that
the axionic fields discussed here need not couple to QCD.
10
This fact might seem to be an irrelevant technicality, but we will see in 5.4.2 that the
mixing (3.64) is the fatal flaw in one otherwise-compelling scenario for inflation in string
theory.
3.2 Compactification 113
11
The structure of shift symmetries arising in the universal hypermultiplet has been stud-
ied in [296298].
114 3 Elements of String Theory
where the integral is taken over the string worldsheet. We recognize (3.72)
as a topological coupling. Expanding BM N (X) around a fiducial point
X(0) 0 yields
12
In this section we follow the conventions of [34].
3.2 Compactification 115
B2 = b (x) . (3.78)
To determine the axion kinetic terms, and hence the decay constants, we
dimensionally reduce the ten-dimensional action for the two-form,
Z Z
1 10 2 1
7 2 0 4 d X |dB2 | d4 x g ( b b ) , (3.79)
2(2) gs ( ) 2
where Z
1
?6 . (3.80)
6(2)7 gs2 (0 )4 X6
f2 1 02
2 . (3.81)
Mpl 6 L4
116 3 Elements of String Theory
Since computational control requires L 0 , we infer that f Mpl .
Qualitatively similar upper bounds on the decay constants occur in all com-
putable limits of string theory that have been explored to date [46, 47].
13
The word moduli actually has several different meanings in different contexts, so a clar-
ification is appropriate. The geometric notion is that moduli parameterize continuous
families of solutions, for example families of Ricci-flat metrics. In physics, a modulus is
a scalar field with gravitational-strength couplings that has vanishing potential at some
level of approximation. Some moduli have exactly vanishing potential before supersym-
metry breaking, while others have vanishing classical potential but obtain a mass from
quantum effects. In some contexts, moduli refers exclusively to parity-even real scalar
fields, as distinguished from pseudoscalar axions, but we will generally refer to complex
moduli.
14
Moduli that decay early, but to fields that themselves linger and affect late-time ob-
servables, are a very interesting exception: see e.g. [303309].
3.3 Moduli Stabilization 117
e8A
2 e4A = |G3 |2 + e4A ||2 + |e4A |2 + 22 e2A Jloc , (3.84)
2Im( )
where 2 is the Laplacian on X6 , and the effects of local sources are pa-
rameterized as
9 3
!
1 X M X M
Jloc T M T M , (3.85)
4
M =4 M =0 loc
with TM N the stress-energy tensor derived from Sloc . In the absence of local
sources, i.e. for Jloc = 0, the solution is trivial, with constant A, constant
, and vanishing G3 . (To see this, note that the l.h.s. of (3.84) integrates
15
For time-dependent solutions, we would require a more general ansatz [310].
118 3 Elements of String Theory
to zero on X6 , while the first three terms on the r.h.s. are all non-negative.)
A non-trivial warped compactification requires one or more sources with
Jloc < 0 [316], for example orientifold planes.
Next, the Bianchi identity for the five-form flux is
dF5 = H3 F3 + 22 T3 loc
3 , (3.86)
where loc3 is the D3-brane charge density due to the local sources. Because
F5 is self-dual, (3.86) may also be thought of as an equation of motion. Inte-
grating (3.86) over X6 leads to a tadpole-cancellation condition (i.e. Gausss
law constraint) Z
1
2 H3 F3 + Qloc
3 =0 , (3.87)
2 T3 X6
where Qloc loc
3 is the total charge associated with 3 . Substituting (3.83) into
16
(3.86) and combining with (3.84), we get
e8A
2 e4A = |iG3 ?6 G3 |2 + e4A |(e4A )|2
24Im( )
+ 22 e2A (Jloc Qloc ) , (3.88)
16
This corrects the numerical factor appearing in [295], cf. [317].
3.3 Moduli Stabilization 119
where I, J run over all the moduli (Ti , and ). Supersymmetry is pre-
served if all F-terms vanish,17 i.e. if
DI W0 I W0 + (I K)W0 = 0 , (3.96)
where I runs over all the moduli.
No-scale structure.The K
ahler potential (3.93) is of a specific form that
satisfies
K0I J I K0 JK0 = 3 .
X
(3.97)
I,J=Ti
17
When gauge multiplets are present in the effective theory, D-term contributions are an
important alternative source of supersymmetry breaking, but our present discussion is
confined to the moduli sector.
3.3 Moduli Stabilization 121
where is a constant, and k(z , z ) is the Kahler potential for the metric on
the Calabi-Yau manifold. The K ahler potential (3.99) is of no-scale type: if
the superpotential W is independent of T and of the z , then the F-terms
of these fields do not appear in the F-term potential. The mixing between
the Kahler modulus T and the D3-brane position moduli implied by (3.99)
has significant ramifications for inflationary model building with D3-branes:
see 5.1.
In summary, in a no-scale compactification with imaginary self-dual
fluxes, one finds, at leading order in 0 and gs , that the vacuum energy
vanishes,18 the complex structure moduli and axiodilaton are stabilized,
the Kahler moduli and D3-brane position moduli have vanishing potential.
Perturbative Corrections
The most famous perturbative correction to the Kahler potential descends
from an (0 )3 curvature correction in ten dimensions, namely the quar-
tic invariant R4 appearing in (2.30). This term is part of the classical,
higher-curvature ten-dimensional supergravity theory: it arises via a four-
loop correction to the -function of the worldsheet -model [198], rather
than from a loop in spacetime. In the four-dimensional effective theory, the
result takes the form [321]
" #
(X6 )(3)
K = 2 ln V + 3/2 , , (3.100)
2gs 2(2)3
where (X6 ) is the Euler characteristic of X6 , and (3) 1.202 is Aperys
constant. The K ahler potential (3.100) does not satisfy the no-scale condi-
tion (3.97) (unless = 0).
Perturbative corrections from loop effects in spacetime, i.e. from higher-
genus string worldsheets, will also generically spoil the no-scale structure
(3.97). The only explicit results available are for N = 1 compactifications
18
Having a non-supersymmetric vacuum with vanishing vacuum energy seems too good
to be true, and it is: no-scale structure on its own is not a solution to the cosmological
constant problem, because it does not survive quantum corrections.
122 3 Elements of String Theory
where i stands for the K ahler modulus associated with the four-cycle
wrapped by the i-th D7-brane. The second term in (3.101) takes the form
3
W
W 1 X E i (, )
K(g = . (3.103)
s) 2
128 i=1 j k
j6=k6=i
Nonperturbative Effects
Although the K ahler potential for the Kahler moduli receives perturbative
corrections in the 0 and gs expansions, the superpotential receives no cor-
rections in either expansion, to any order in perturbation theory, as we now
explain.
The fact that the superpotential of a supersymmetric field theory receives
no perturbative corrections in the ordinary ~ expansion, corresponding to
the gs expansion in string theory, was originally established directly [326].
Elegant non-renormalization theorems in string theory [299302] arrived at
the same end by combining holomorphy and shift symmetry arguments.
In the heterotic string setting emphasized in [299302], the argument for
non-renormalization in gs is more straightforward than in type IIB flux
3.3 Moduli Stabilization 123
4
Z Z
1 4A(y)
d4 x g Tr F F ,
S= 2 d gind e (3.104)
2g7 4
where the indices are raised with the unwarped metric g , and g7 is the
gauge coupling of the (7+1)-dimensional Yang-Mills theory,
and gind is the induced metric on the D7-brane. Because of the appearance
of e4A(y) , V4 as defined in (3.107) is sometimes called the warped volume.
Given certain topological conditions on 4 , discussed further below
heuristically, one asks that 4 have no deformations that could correspond
to charged matter fields the four-dimensional gauge theory arising upon
dimensional reduction is pure glue N = 1 super Yang-Mills theory. At low
124 3 Elements of String Theory
19
Supersymmetry requires that the superpotential is a holomorphic function of the moduli,
but verifying that V4 is the real part of a holomorphic function is highly non-trivial
[218, 335]. When D3-branes are present, their backreaction on the volume V4 must be
incorporated in order to maintain holomorphy [335]. This effect was first understood in
the open string channel, as a threshold correction to the gauge coupling g [336].
20
An elliptic fibration is a fibration in which almost all fibers are non-singular and have
the topology of two-tori, but a finite number of singular fibers can appear. The possible
singular fibers have been classified by Kodaira in [339].
3.3 Moduli Stabilization 125
F-theory [340]: one says that F-theory has been compactified on Y , which
is an elliptically-fibered four-fold over the base X.
Witten observed in [337] that a necessary condition for a non-vanishing
Euclidean D3-brane superpotential term associated with a four-cycle 4
X is that 4 is the projection of a six-cycle D Y obeying
3
(1)i h0,i (D) = 1 .
X
(OD ) (3.111)
i=0
where OD denotes the trivial line bundle defined on D (see e.g. [341] for
the relevant mathematical background). The number (OD ) is known as
the holomorphic Euler characteristic of D [341], or the arithmetic genus of
D.21 Next, a sufficient condition for a non-vanishing Euclidean D3-brane
superpotential is [337]
h0,1 (D) = h0,2 (D) = h0,3 (D) = 0 . (3.112)
A six-cycle D obeying (3.112) is said to be rigid: the Hodge numbers in
(3.112) count the independent deformations of D.
The sufficient condition (3.112) is unmodified by the presence of flux,
but in flux backgrounds the necessary condition (3.111) is modified and
becomes less restrictive [334, 343348]. Couplings to flux can give mass
to (some of the) deformations of Euclidean D3-branes, and of D7-branes,
counted by h0,2 . Generalizations of (3.111) to backgrounds with flux, and
further consistency conditions, are described in [334, 343350] and reviewed
in [338, 351].
21
In the mathematics literature, some authors define the arithmetic genus pa (D) so that
pa (D) = 1 (OD ), for D a six-manifold [341]. Here, as in most of the string theory
literature, the arithmetic genus and the holomorphic Euler characteristic are both equal
to (OD ), cf. [337, 342]. In (3.111), the notation (OD ) for the holomorphic Euler
characteristic is used instead of (D) because the latter can be confused with the more
familiar topological Euler characteristic.
126 3 Elements of String Theory
and when they are computable, they are not important [351]. To understand
the observation of [352] in more detail, let be a modulus that controls a
weak coupling expansion, such that is the free limit. Concretely,
could be the K ahler modulus that measures the compactification volume,
= T + T, so that corresponds to decompactification to ten di-
mensions; or, for the string loop expansion, = gs1 = e . We now ask
whether perturbative or nonperturbative corrections generate a potential
for that has a minimum at finite . Because the leading-order classical
action is valid for , the potential V () generated by perturbative and
nonperturbative corrections must vanish for . In particular, V ()
must approach zero from above or from below as (see fig. 3.2). If
V () is positive for large , then the leading correction term in V (), which
dominates for , creates an instability that drives the theory toward
= . If instead V () is negative for large , then the leading correc-
tion to the free theory creates an instability that drives the theory toward
smaller , and hence toward stronger coupling. Either way, the leading
correction term creates an instability, and a (meta)stable vacuum can arise
only if higher-order corrections make comparably important contributions
that counterbalance the instability. But once two22 consecutive terms in the
weak coupling expansion are comparable, one expects that the entire series
must be included. While it could happen that the first and second non-
vanishing terms are competitive because the second is accidentally large,
verifying that this leads to a consistent solution requires examining higher
terms in the series to rule out unanticipated accidental enhancements at
higher orders. Thus, metastable vacua are quite generally found at points
in moduli space where the weak coupling expansions break down. This fact
presents a major obstacle to the search for metastable string vacua, be-
cause in nearly all cases, at most the first non-vanishing correction in each
expansion (0 or gs ) is known explicitly.
Fig. 3.2. The Dine-Seiberg problem [352] for a modulus . In case (a), there is
a runaway to = , where the theory is free. In case (b), the leading correction
drives the theory toward small , where it is strongly coupled. The existence of
the minimum in case (c) requires competition among at least three terms.
22
When V > 0 for large , three separate terms are required see [351, 353].
3.3 Moduli Stabilization 127
KKLT Scenario
The seminal KKLT proposal [356] for constructing stabilized vacua bypasses
all perturbative corrections and instead makes use of nonperturbative con-
tributions to the superpotential.
In the presence of three-form flux the complex structure moduli and dila-
ton acquire supersymmetric masses via the classical superpotential (3.94),
cf. 3.3.1. If we denote the typical mass scale by mflux , then at energies
E mflux the complex structure moduli and dilaton can be integrated out
(see the discussion below), and the classical superpotential W0 becomes a
constant. The fields remaining in the low-energy effective theory are the
Kahler moduli,24 which do not appear in the classical superpotential.
23 0
Whether the leading perturbative correction to the potential comes from the first
correction to K, or instead from the first gs correction to K, is not obvious a priori,
and can depend on parameter values see 5.5 for a detailed discussion.
24
If spacetime-filling D3-branes are present, their positions are also light fields in the
128 3 Elements of String Theory
1.0
dS 150 200
100 250 300
-1.0
SUSY-AdS
-2.0
This potential is plotted in fig. 3.3 (dashed line). It is easy to see that the
vacuum solution is supersymmetric anti-de Sitter space. Letting (T + T)?
be the value of the Kahler modulus at the minimum, we find (T V(np) )? =
(DT W )? = 0 and
a(T +T)? 2
W0 = A e 1 + a(T + T )? . (3.116)
3
Control over the instanton expansion of the superpotential, corresponding
to neglecting the ellipses in (3.113), requires that a(T + T)? 1. Moreover,
perturbative (0 and gs ) corrections to the Kahler potential (3.93) may
be neglected if (T + T)? 1.26 We see from (3.116) that the volume is
stabilized in a controlled limit only for an exponentially small value of the
flux superpotential, W0 A. This can be achieved through a fine-tuned
choice of quantized flux, following [193].
A number of authors have critically examined the two-step procedure of
integrating out the complex structure moduli and dilaton, and then studying
the effective theory for the Kahler moduli, instead of analyzing all moduli on
the same footing [361365]. The underlying justification for a two-step pro-
cedure is that the mass scale mflux is set by the flux quantization condition,
and does not diminish as W0 is fine-tuned to be small, whereas the mass of
the Kahler modulus T at the minimum is proportional to W0 . To under-
stand this, we expand the flux superpotential around the supersymmetric
minimum,27
26
String loop corrections to K are suppressed at large volume, and not only by factors of
gs , because K0 involves V, so that any additive correction to K is subleading in volume;
see 5.5.
27
To be precise, we mean the minimum determined by D W0 = D W0 = 0, where we
stress that W0 is the flux superpotential (3.116), not the full superpotential (3.113).
130 3 Elements of String Theory
2 + 7V + V2
K 2 3 2 1
V(0 ) = 3 e 2 W0 W0 3 , (3.118)
4
V 2V + V
28
Some care is needed to ensure that holomorphy is maintained in this process, as ex-
plained in [362] and described in an explicit example in [364].
3.3 Moduli Stabilization 131
one finds29
(
h
V(np) + V(0 ) = eK K j aj Aj a A e(aj Tj +ai Ti )
K + a A eai Ti W j K
i
aj Aj eaj Tj W
)
3 2 1
+ W0 . (3.119)
4 V
At very large volume, the perturbative term (3.118) dominates over the
nonperturbative terms (3.114). Competition between (3.118) and (3.114)
can occur if one or more cycles are exponentially smaller than the largest
cycles. Denoting the small cycle volumes by s 12 (Ts + Ts ), the idea is to
take the limit
V , with as s = ln V . (3.120)
Along the ray in the K ahler moduli space defined by (3.120), the exponen-
as Ts
tials e in (3.114) are proportional to 1/V, and all terms in (3.119) are
of the same order in 1/V. Notice that the hierarchy (3.120) is only possible
for h1,1 1,1
+ > 1 we will therefore take h+ > 1 for the remainder of this
discussion.
The sign of is determined by the topology of the compactification, with
> 0 corresponding to (X6 ) < 0. In this section we will assume that > 0,
which implies that the contribution (3.118) is positive, so that the poten-
tial (3.119) approaches zero from below at large V along the ray (3.120). To
establish the existence of a minimum, one then needs to argue, first, that
the potential along (3.120) becomes positive at sufficiently small V, so that
by continuity the potential restricted to (3.120) is minimized at an interme-
diate point V? . Second, one must show that at V? , (3.119) is non-decreasing
in the h1,1
+ 1 directions in the K ahler moduli space that are perpendicular
to the ray (3.120).
A useful heuristic argument that is valid in certain simple cases (with
provisos enumerated below) goes as follows. If the term (3.118) is dominant
over the exponential terms at small volume, this establishes that (3.119)
likewise becomes positive at small volume. Next, if the leading exponential
terms in (3.119) are positive, and all h1,1 + 1 K ahler moduli appear in
the nonperturbative superpotential, the potential increases in the directions
transverse to the ray (3.120). In combination, these assumptions imply the
existence of a minimum at exponentially large volume. This minimum has
29 0
For the and string loop expansions to be valid, we require V 1, as discussed
0
further below; see [366] for a systematic exposition of the expansion in this setting.
132 3 Elements of String Theory
To check for the existence of a minimum in the limit (3.122), one needs to
examine in detail the inverse K ahler metric K i , and in particular the block
corresponding to the small cycle moduli Ts(a) . A systematic treatment for
Ns = 1 and Ns = 2 appears in [367].
To understand the results of [367], one piece of geometrical background
is necessary. Suppose that M is a complex manifold (potentially containing
singularities) of complex dimension n, and let p be a point in M. The blowup
of M at a non-singular point p replaces p with a copy of Pn1 , known as the
exceptional divisor. The blowup of a singular point can result in more gen-
eral exceptional divisors. When the blowup of M is a Calabi-Yau threefold,
the exceptional divisor is a four-cycle, with size parameterized by one of the
Kahler moduli. When the exceptional divisor satisfies the rigidity condi-
tion (3.112), the corresponding Euclidean D3-brane superpotential term is
non-vanishing [337].
A necessary condition for an LVS minimum is that at least one of the Ns
1 small cycles is a rigid exceptional divisor arising from blowing up a singular
point [367]. When Ns = 1, this condition guarantees that (3.119), restricted
to the ray (3.122), has a minimum at exponentially large volume. Whether
this is a minimum of the full potential depends on the curvature in the Nb 1
directions perpendicular to (3.122), as we discuss further below. For the case
Ns = 2, if the two small cycles correspond to blowups of distinct points,
then (3.119) restricted to (3.122) again has a minimum at exponentially
large volume, with the same caveat about transverse directions. If instead
the two small cycles are two independent resolutions of the same singular
point, then an LVS minimum along (3.122) exists only if there is a basis in
which the volume V is symmetric in the two Kahler moduli Ts1 and Ts2 . For
a discussion of the necessary conditions on K i in the context of a survey of
a class of Calabi-Yau manifolds, see [368].
A canonical class of examples of LVS vacua arise in what are known as
3.3 Moduli Stabilization 133
with r > 0. This compactification has a single large four-cycle, with vol-
ume b , and Ns = h1,1 r
+ 1 small four-cycles, with volumes s . Increasing
r
one of the s with all else fixed decreases V, so the small cycles act like
holes in a large cheese. The structure (3.124) can arise if the Ns small cy-
cles correspond to the blowups of Ns distinct singular points. In the case
of a compactification of strong Swiss cheese form (3.124), the necessary
conditions described in [367] are readily met, for any Ns > 0.
The final, critical question is whether the potential is stable or unstable in
the Nb 1 directions perpendicular to (3.122). In fact, (3.119) per se, which
includes only the leading 0 correction to K, viz. (3.100), has Nb 1 flat
directions. The exact moduli potential, incorporating all perturbative and
nonperturbative effects in gs and 0 , very plausibly depends on the Nb 1
fields that are unlifted by (3.119).31 However, appealing to an unknown and
uncomputable potential to lift these remaining moduli is problematic, not
least because there is no evidence that the resulting masses-squared will all
be positive. That is, further perturbative corrections beyond (3.100) could
well introduce instabilities along one or more of the Nb 1 flat directions of
(3.119), leading to an LVS saddle point rather than a minimum.32 Indeed,
as we argue in 3.5.3 below, in certain ensembles of supergravity theories it
is overwhelmingly improbable that all Nb 1 flat directions are stabilized
rather than destabilized: the probability of stability is exponentially small
in Nb . Whether the assumptions of 3.5.3 are applicable to the moduli
potential in LVS is an important open question (see [369] for recent work).
30
See [368] for a study of the incidence of the form (3.123) in a class of Calabi-Yau
manifolds.
31
As we will explain in 5.5.2, it has been suggested [43,367] that the leading gs correction
to K, with form conjectured in [324] following computations in [322, 323], can stabilize
the Nb 1 flat directions. However, a more detailed demonstration of stability would
be valuable.
32
However, if it can be established that the potential increases as one moves toward each
of the boundaries of the moduli space, then one can again make a continuity argument
for the existence of a minimum. We thank Joe Conlon and Fernando Quevedo for
discussions of this point.
134 3 Elements of String Theory
In summary, the necessary conditions for an LVS minimum are the fol-
lowing: > 0; h1,1
+ Ns + Nb > 1; Ns 1 K ahler moduli corresponding
to the blowups of points. For Ns > 1, further conditions on the blowups
are necessary [367], while for Nb > 1, it is necessary that further correc-
tions, beyond (3.119), render stable the Nb 1 flat directions of (3.119).
Explicit examples with Nb = 1 that meet all other necessary criteria are
now well-known [357, 366, 367].
Several differences between LVS and the KKLT scenario should be em-
phasized. In LVS, some cycles are exponentially larger than others, while in
KKLT the cycles are not hierarchically different in size. In KKLT, the clas-
sical flux superpotential W0 is fine-tuned to be exponentially small, while
in LVS W0 is of order unity. In KKLT, the AdS4 vacuum is supersymmet-
ric, whereas in LVS the AdS4 vacuum is non-supersymmetric. However,
in both scenarios some form of uplifting effect is required to achieve a de
Sitter vacuum, as we now explain.
for all fields, must of course be solved simultaneously. This presents a dif-
ficulty, because the vacuum energy contribution from the uplifting sector
cannot be a perturbatively small correction to the original vacuum energy.
In most approaches the stabilization in AdS is analyzed in a supersymmet-
ric effective action, and one must take care that the large supersymmetry
breaking from the uplifting sector does not invalidate this treatment. In
summary, the task in uplifting is to identify a sector that breaks super-
symmetry dynamically, in a parametrically controlled manner, and makes a
positive contribution to the vacuum energy without disrupting the physics
that led to a stabilized AdS vacuum. As we explain in the following, these
requirements are very challenging, even taken in isolation.
First of all, one must engineer a sector of fields that breaks supersymme-
try. As a concrete example, consider placing multiple D-branes at the sin-
gular apex of a Calabi-Yau cone, leading to a supersymmetric gauge theory
in four dimensions. Some of the resulting gauge theories have metastable
vacua in which supersymmetry is dynamically broken [370376], while in
other cases, such as [377379], there are runaway instabilities in directions
parameterized by K ahler moduli [378, 380, 381]. But even if one finds a
configuration of D-branes on a noncompact Calabi-Yau cone leading to a
flat space gauge theory that dynamically breaks supersymmetry, establish-
ing that metastability survives compactification is highly non-trivial (but
see [374]). The essential issue is that in the low-energy Lagrangian of a
compactification, all parameters are determined by the vevs of fields, and
are therefore dynamical at sufficiently high energies. Any gauge theory con-
struction relying on a non-dynamical parameter for example, the mass
of a quark flavor, as in [382] is potentially vulnerable, upon compactifi-
cation, to an instability along which this parameter evolves. Often a second
stage of model-building is required in which one generates the desired vev
dynamically and establishes the absence of instabilities see e.g. [373].
After identifying a supersymmetry-breaking sector, one must compute the
effects of supersymmetry breaking on the remaining fields. A pervasive but
potentially deceptive picture for uplifting is that the uplifting sector exists
somewhere else in the compactification: stabilization in AdS is imagined to
result from sources and fields in one region, while supersymmetry breaking
arises in another region, and the vacuum energy contributions are therefore
approximately additive, by locality in the extra dimensions. One problem
with this modular picture, as we explain in detail in 4.2, is that geometric
separation does not imply complete decoupling of two sectors. At the very
least, the supersymmetry-breaking sector interacts with the remaining fields
by its coupling to the overall compactification volume V: any source S of
positive energy33 in the four-dimensional theory must be negligible in the
33
A ten-dimensional cosmological constant would be an exception, but this is excluded
136 3 Elements of String Theory
by ten-dimensional supersymmetry.
3.4 De Sitter Vacua 137
34
The two-derivative qualifier refers to omission of higher-curvature contributions, and
is usually assumed implicitly.
138 3 Elements of String Theory
lar near the source. It has been argued in [435] that the singularity
in the flux is not a consequence of linearization: the nonlinearly back-
reacted, but still smeared, solution displays singularities. This leaves
smearing as perhaps the most plausible cause of the singularities. (See
e.g. [433,438,444] for related work on the problem of singularities from
localized sources.)
. Smearing and brane polarization.What sort of smooth supergrav-
ity solution might one expect for p gs1 non-smeared anti-D3-
branes? As noted in [425], anti-D3-branes that are initially coincident
are driven to redistribute themselves along an S 2 in the S 3 , mani-
festly breaking some of the symmetries preserved by a configuration
smeared on the S 3 . This process can be viewed as polarization of the
branes [288] by the flux background, as in the related solution found
by Polchinski and Strassler [445], where brane polarization resolves
the singularity present in the unpolarized configuration. In [429], it
was conjectured that a smooth anti-D3-brane solution can be modelled
on the system in [445], with polarization of the anti-D3-branes along
an S 2 S 3 being responsible for removing the singularities.35 Such a
solution is clearly incompatible with a smearing approximation, but
solving the equations of motion in this setting is a formidable technical
challenge, and at present it is not known whether brane polarization
will resolve the singularities.
35
D-branes can generally polarize in multiple ways, and an alternative to the polarization
2 3
identified in [425], where the anti-D3-branes spread along an S S , is for the anti-
2 2
D3-branes to spread along a different S , namely the S that shrinks toward the tip of
the throat (see 5.1.1). This process moves the anti-D3-branes radially outward, away
from the tip, a direction of motion that is opposed by the classical potential from fluxes.
It was shown in [436] that this alternative, radial polarization is not possible, but this
does not exclude the expected polarization of [425].
140 3 Elements of String Theory
electric field. Of course, the divergence in the energy of the electric field is
removed in the quantum theory.
The question, then, is whether the singularities seen in [427, 429, 430] are
expected, and therefore plausibly resolved in the exact solution. A central
concern raised by [427] is that the singularities in three-form flux do not
appear to have a distinct physical origin [427]. That is, according to [427]
it is not obvious how the anti-D3-brane can serve as a source for singular
three-form flux, and correspondingly these singularities are unexpected.
It is certainly true that the only flux sourced by an anti-D3-brane in empty
flat space is five-form flux F5 , just as for a D3-brane: see the Chern-Simons
coupling eq. (3.22). For an anti-D3-brane in a classical flux background, the
problem is more subtle: the supergravity equations of motion are nonlinear,
and the various fluxes are coupled to each other, as we will explain below. To
understand this case, we begin by developing intuition in a simpler example.
Let us see how a point source of one field A, in a classical background of
a second field B, can source a singular profile of a third field C, even if the
source does not have a direct coupling to C in the Lagrangian. Consider
classical four-dimensional electromagnetism coupled to an axion , with
Lagrange density
1 1
L = ()2 F F F F , (3.130)
2 4 f
where f is the axion decay constant. Suppose that there is a constant
classical background magnetic field B = B z , and place an electron in this
field, at rest at the origin. The electron sources an electric field E = e r /r2 ,
so that in spherical polar coordinates (r, , ) one has
eB
EB = cos . (3.131)
r2
The equation of motion for is therefore
eB
2 = cos . (3.132)
f r2
Thus, the axion effectively has a local source, even though the electron
alone does not couple to . This example illustrates that in theories with
local sources and multiple coupled fields, not every singular field profile
arises from a one field, one source coupling in the Lagrangian: classical
background fields can also play a role.
In the case of an anti-D3-brane in a KS throat, the classical background
field analogous to B is the three-form flux G3 of the KS solution. Schemat-
ically, the anti-D3-brane sources singular five-form flux, which couples to
the non-singular background three-form flux, and thereby sources a singu-
lar three-form flux. To see this more explicitly, we consider the equation of
3.5 Statistics of String Vacua 141
36
The four-dimensional spacetime is often assumed to be maximally symmetric, i.e. de
Sitter space, Minkowski space, or anti-de Sitter space. Isolated solutions without four
large spacetime dimensions could also be considered to be part of the landscape, but
we will focus on the class of vacua that are directly relevant for cosmology.
3.5 Statistics of String Vacua 143
37
Limitations and weaknesses of the current evidence have been described in e.g. [427,
436, 437, 451453].
144 3 Elements of String Theory
In practice, dNmin is far more challenging to study than the related index
density dImin , defined by
( i ) (1)Fi ,
X
dImin () (3.136)
i
where (1)Fi is the sign of the determinant of the fermion mass matrix
(see [44]). The integral of dImin over the moduli space is manifestly not the
total number of vacua: it is instead a sum weighted by signs. The advantage
of considering dImin is that it is computable: one can obtain the elegant
Ashok-Douglas formula [454]39
X (2Lmax )b3
dImin = det(R ) , (3.137)
LLmax b3 /2 b3 !
where is the K ahler form on the moduli space, R is the curvature two-
form, b3 is the third Betti number of the compactification, and the number
Lmax represents a tadpole constraint on the flux. Equipped with (3.137), one
can estimate the actual number of vacua by attempting to place bounds on
the degree of difference between dImin and dNmin . One pivotal observation
is that the number of vacua is exponential in b3 .
38
The F-terms described here are those due to the classical flux superpotential W0 , but
nonperturbative contributions to the superpotential for example, from Euclidean
D3-branes introduce further dependence on the complex structure moduli.
39
Evidence supporting the result (3.137) in explicit flux compactifications on Calabi-Yau
three-folds was obtained in [455, 456], building on [457].
3.5 Statistics of String Vacua 145
There are two critical caveats that prevent one from concluding at this
stage that type IIB string theory compactified on a Calabi-Yau manifold
with large b3 has an exponentially large number of metastable de Sitter
vacua. First, we have thus far described only the complex structure moduli,
and a local minimum of the potential on MC may or may not correspond to a
local minimum of the exact potential on the full moduli space Mtotal , which
also includes the Kahler moduli and the positions of D-branes. Second, MC
is noncompact, as is Mtotal : in particular, the Kahler moduli space MK can
be continued toward infinite volume, where one recovers ten-dimensional flat
space. Noncompactness of MC implies that VF may not have a minimum in
MC .40 Thus, one is not strictly guaranteed any vacua for a given choice of
flux. Equation counting does certainly suggest that VF will generically have
one or more minima inside MC , but topology does not necessitate this.
With this background, we emphasize that the celebrated counting of 10500
vacua in the landscape (cf. [44]) does not refer to a counting of metastable
vacua of the full potential for all moduli (at any level of approximation): it is
a counting of supersymmetric vacua of the complex structure moduli sector,
neglecting the K ahler moduli and postponing the question of metastable
supersymmetry breaking.
Let us therefore ask whether one can extrapolate from this result to esti-
mate the number of de Sitter vacua in type IIB flux compactifications. One
might be tempted to argue as follows: suppose that one single metastable
de Sitter vacuum is found, e.g. a KKLT solution on a particular Calabi-
Yau with a particular choice F? of quantized flux. As famously explained
by Bousso and Polchinski [193], the many possible choices of quantized p-
form flux in compactifications with many p-cycles lead to a discretuum of
closely-spaced vacuum energy densities. Can one then apply this logic and
appeal to the existence of many fluxes F0? , F00? , that differ (by discrete
quanta) from F? , but lead to a very similar cosmological constant, in order
to replicate the single de Sitter vacuum into O(NF ) de Sitter vacua? No:
the fact that VF? has a metastable local minimum in no way implies that
VF0 has a local minimum. This fact can also be understood in concrete
?
examples: a change of quantized fluxes that leads to a small change in the
cosmological constant generally involves large changes in the individual flux
quanta, and correspondingly makes an order-unity change to the effective
action, entirely changing the distribution of extrema (if any exist).
One must therefore be cautious when using the vast number of super-
symmetric (or no-scale supersymmetry-breaking) vacua in the complex
structure moduli sector, cf. (3.137), to argue for the existence of a compa-
rable number of metastable de Sitter vacua of the full potential on the total
40
For MK , one manifestation of the corresponding fact is that the potential can have its
minimum at infinite compactification volume.
146 3 Elements of String Theory
moduli space: NdS 6= NF in general. We will now discuss this issue in detail.
41
In some circumstances one can show that the number of critical points per choice of
flux is exponentially large. We thank Edward Witten for this observation.
3.5 Statistics of String Vacua 147
V = eK Fa F a 3|W |2 .
(3.142)
The object of primary interest is the Hessian matrix H at a critical point p
of the potential,
2 2
ab V ab V
H= . (3.143)
a2b V a2b V
At a local minimum of the potential, the eigenvalues 1 2 . . . N of H
are nonnegative, so
fmin = P (1 > 0) , (3.144)
where as explained above, the probability P is computed in the ensemble
consisting of the Hessian matrices at each of the critical points in C.
To express H in a convenient form [48, 458], we perform a coordinate
transformation to set Kab = ab at p, and a Kahler transformation to set
K = 0 at p. We denote the geometrically-covariant and Kahler-covariant
derivative by Da , and define the first three covariant derivatives of the su-
perpotential as
Fa Da W , Zab Da Db W , Uabc Da Db Dc W . (3.145)
The Hessian then takes the form [48, 458]
!
Zac Zbc Fa Fb RabcdF c F d Uabc F c Zab W
H=
U abc F c Zab W Zac Zbc Fb Fa RbacdF c F d
+ 1 F 2 2|W |2 , (3.146)
where indices are raised with ab , 1 is the 2N 2N identity matrix, and
Rabcd is the Riemann tensor of the metric on field space.
The idea at this stage is to recognize that the large dimension N of
the field space need not remain an obstacle, but can instead be an expan-
sion parameter! The Hessian is a large matrix, and random matrix theory
[459] provides a powerful tool for determining its eigenvalue spectrum. The
foundational insight in random matrix theory [460] is that one can make
sharp predictions about the statistical properties of the eigenvalues of a
large (N N ) diagonalizable matrix given very limited information about
the actual entries of the matrix. The guiding principle here is universality,
which states that for N 1, the statistics of the eigenvalues have little
dependence on the statistics of the matrix entries. This may be thought of
as central limit behavior for matrices.
148 3 Elements of String Theory
42
The fine print is that the correlations cannot be too numerous [464], and the statistical
distributions must have appropriately bounded moments. Universality has been formu-
lated and established rigorously in many settings cf. [461, 465] but to simplify the
discussion we will continue to omit the associated technicalities. More details can be
found in [48].
3.5 Statistics of String Vacua 149
then (3.148) holds for a typical member of the ensemble, while approximate
supersymmetry as in (3.147) can occur via a rare fluctuation. We will refer
to the ensemble of critical points generated via (3.149) as generic critical
points; it was argued in [48] that the overwhelming majority of critical
points are in fact of this form.
In a random supergravity theory with N chiral superfields, the Hessian
(3.146) is a 2N 2N matrix whose entries are stochastic variables. Con-
siderable structure is evident in (3.146), and the next step, following [48],
is to decompose (3.146) into a sum of constituent random matrices with
simple properties. The eigenvalue spectrum of H which is the quan-
tity controlling the probability of metastability can then be obtained by
appropriately convolving the spectra of the constituents.
To this end, we briefly outline the properties of two classic ensembles
of random matrices. The (complex) Wigner ensemble, also known as the
Gaussian Unitary Ensemble, consists of N N Hermitian matrices M of
the form
M = A + A , (3.150)
where the entries Aij are stochastic complex variables with uniformly-distri-
buted phase and normally-distributed magnitude, |Aij | N (0, ). The
eigenvalue density () of a typical member of the Wigner ensemble is given
by the Wigner semicircle law,
q
1
() = 4N 2 2 . (3.151)
2N 2
Next, the complex Wishart ensemble consists of matrices of the form
M = AA , (3.152)
with N 2 (1 P/N )2 .
p
We are now prepared to use random matrix theory to analyze the eigen-
value spectrum of the supergravity Hessian matrix (3.146), in a random
supergravity theory. We will begin by studying generic critical points, as
defined by (3.149). One recognizes (3.146) as the sum of constituent matri-
150 3 Elements of String Theory
Zac Zbc
0
HZ (3.154)
0 Zac Zbc
An analytic expression for (HW W W ) was obtained in [48] (we omit it here
for brevity). In fig. 3.4 we compare a histogram of the eigenvalues of the full
Hessian matrix (3.146) in random supergravity, making no approximation,
to the analytic result of the Wigner Wishart Wishart (W W W ) model
(3.156). The model has no freely-adjustable parameters: we take N = 200
fields in both the simulations and the analytic model. The agreement is
excellent; the slight tail at the right edge is a consequence of finite N .
Although formal results in the subject often require the limit N ,
fig. 3.4 makes it clear that N = 200, which is an entirely reasonable number
of fields in Calabi-Yau compactification, is a sufficiently large value of N .
We conclude that the Hessian matrix (3.146) at a generic critical point of
a random supergravity theory is very well approximated by the analytic
model (3.156).
At first glance, the eigenvalue spectrum (3.156) depicted in fig. 3.4 may
appear to determine fmin as follows: according to the probability density
(), a single eigenvalue is positive with probability
R
0 ()
f> R , (3.157)
()
suggesting that fmin = (f> )N . This is not correct: () describes the eigen-
value density for a typical Hessian matrix in the ensemble, and can be used
to compute the probability that a single eigenvalue i falls in some given
interval, provided that the remaining N 1 eigenvalues are unconstrained.
However, the eigenvalues of a random matrix are strongly correlated, and
3.5 Statistics of String Vacua 151
0 2 4 6 8
Fig. 3.4. The histogram shows the spectrum of eigenvalues of the full Hessian
matrix in random supergravity (for N = 200 fields), while the curve gives the
analytic result from the W W W model (3.156) [48]. The curve is a parameter-free
prediction of the model, not a fit. (Figure adapted from [48].)
43
The extreme scarcity of minima in a landscape whose Hessian matrices are governed by
152 3 Elements of String Theory
Wigners Gaussian Orthogonal Ensemble was first discussed in the context of cosmology
in [473]. An analysis of uplifting of supersymmetric AdS vacua of type IIA string theory,
leading to the same conclusion, appears in [474].
44
Unbroken supersymmetry in AdS does not guarantee the absence of tachyons allowed
by the Breitenlohner-Freedman bound [475], but the techniques described above can be
used to compute the probability that there are no tachyons [476]. See also [477].
45
The principal instability corresponds to the scalar partner of the Goldstino see [478]
for an analysis of geometric conditions that ensure stability in this direction, and [479,
480] for discussions of inflation along this direction.
46
A second important example is the Large Volume Scenario, cf. 3.3.3, which leads to
non-supersymmetric AdS4 vacua that have been argued to be automatically tachyon-
free in certain cases [357, 367]. The incidence of instabilities in LVS has been analyzed
in [369].
3.5 Statistics of String Vacua 153
47
See also [482], where the probability of metastability in a Gaussian landscape was shown
to be anticorrelated with the magnitude of the vacuum energy.
4
What is String Inflation?
154
4.1 From Strings to an Inflaton 155
I II
UV-completion
spontaneously
broken SUSY
fine-tuning
or symmetry
1
Lighter moduli, with m H, may be natural in certain circumstances: see e.g. [489].
2
The primary examples are Dp-branes with p 3, NS5-branes, or M5-branes, wrapping
suitable cycles. Orientifold planes, in contrast, are non-dynamical: their positions are
not parameterized by light scalars.
158 4 What is String Inflation?
4.1.4 Approximations
In an ideal world, one would derive the inflaton action from first principles,
beginning with fundamental integer data C for a compactification, solving
the equations of motion of the ten-dimensional effective supergravity theory
S10 , order by order in 0 and in gs , and then integrating out massive degrees
of freedom (including the Kaluza-Klein modes of the compactification) to
arrive at a four-dimensional effective theory S4 .3 Unfortunately, computing
the effective action in a metastable non-supersymmetric compactification
is a formidable technical challenge: a direct approach, making no approxi-
mations, would require unforeseen advances in our understanding of string
theory. Indeed, even in compactifications that preserve N = 2 supersym-
metry in four dimensions, such as compactifications of type II string theory
on Calabi-Yau three-folds, the metric on the internal space cannot be com-
puted analytically; in non-supersymmetric solutions, the difficulties are far
greater. In practice, a four-dimensional effective theory is deduced based
on a partial specification of the compactification data C, and an arsenal of
approximation schemes is used in place of a complete calculation.
We will outline some of the most important expansion parameters, sys-
tematic (and non-systematic!) approximation schemes, and simplifying as-
sumptions that are used in determining four-dimensional effective theories
and extracting their dynamics. It will be important to remember these
limitations when we present our case studies in Chapter 5.
. 0 expansion.The 0 expansion is reliable when the gradients of the
background fields are small in units of 0 . However, the compacti-
fication volume is finite, and is typically restricted by the desire to
achieve the hierarchy H < MKK , so that inflation is inherently four-
dimensional. As a result, the 0 expansion is often a barely-controlled
approximation scheme, rather than a convergent parametric expan-
sion, in the regions of interest. This issue is particularly severe in
models of high-scale (equivalently, large-field) inflation, because the
lower limits on the Kaluza-Klein mass become more stringent.
. String loop expansion.The weak coupling approximation retains only
the leading terms in gs 1. One might hope to make gs = e small
3 0
More ambitiously, one might pursue solutions that are exact in and gs , but there has
been very little progress in this direction.
160 4 What is String Inflation?
4
In many settings it is more efficient to compute the couplings between two separated sec-
tors by working in supergravity, rather than by integrating out stretched open strings.
In this closed string approach, one finds a supergravity solution that incorporates the
backreaction of sector A, and then evaluates the probe action for sector B at the appro-
priate location in this solution, in order to determine the effect of sector A on sector B.
This method has been used, for example, to determine the coupling between D3-branes
and quantum effects stabilizing the K ahler moduli [335].
5
If B is taken to be a supersymmetry-breaking sector, and A is the visible sector, then
the notion of decoupling described here corresponds to what is called sequestering [508]
in the literature on supersymmetry breaking. Investigations of sequestering in string
theory [509514] have confirmed that complete decoupling is extremely rare, but partial
suppression of some couplings can occur in certain cases [512].
164 4 What is String Inflation?
Thus, MAB /Mpl < 1 when the volume is controllably large: the stretched
string mass cannot parametrically exceed the Planck mass in an isotropic
compactification. Consequently, the couplings between spatially separated
D-brane sectors will generically be at least gravitational in strength: the
corresponding operators will be suppressed by no more than the Planck
mass.
Isotropy is a strong assumption, and it is important to check whether
decoupling arises automatically in suitably anisotropic compactifications. If
the compactification has p large directions of size L and 6p small directions
of size S, then
1 p1 1 (6p)
MAB ` 2 `s 2
. gs s , (4.7)
Mpl L S
so that for p > 1 the coupling is again at least gravitational in strength at
large volume. (For the case p = 1, see 4.3.) A significant example consists
of a warped throat geometry: a warped cone over an angular manifold X5 is
an example of a highly anisotropic space, if X5 is chosen appropriately for
example, one might consider X5 = S 5 /Zk for k 1. Rather surprisingly, it
was shown in [243] that for any X5 , the stretched string mass is less than
the Planck mass (see 5.1.1). Thus, slender warped throats do not evade
the general argument that gravitational-strength couplings are unavoidable.
The fact that compactness prevents decoupling leads to important con-
straints on the interactions between localized sources. We will illustrate
the issues in the example of a D3-brane/anti-D3-brane pair in a general
unwarped six-manifold X6 , though the problem is more general (see 5.1).
The Coulomb potential of a D3-brane/anti-D3-brane pair separated by a
distance r is !
1 T3 gs2 2
V (r) = 2T3 1 3 , (4.8)
2 r4
where T3 is the D3-brane tension (3.28), and is the gravitational coupling
defined in (3.11). The canonically-normalized field is related to r by
6
We display only the parametric scaling: factors of 2 can be important for this relation,
but depend on the precise geometry and must be analyzed on a case-by-case basis.
4.2 The Eta Problem 165
= T3 r. Computing the slow-roll parameter , we find
10 V
, (4.9)
3 r6
where we have used (3.47) and T32 gs2 2 = . The Coulomb potential (4.8) is
evidently steep at small separations and grows flatter at large separations.
However, the brane-antibrane pair cannot be separated by a distance greater
than the diameter of the compactification, so unless X6 is highly anisotropic,
the potential (4.8) is too steep to support inflation [515].7 This is one of the
simplest examples of the phenomenon of non-decoupling described above.
(y y 0 ) 1
2y0 G(y; y 0 ) = 2y G(y; y 0 ) = + . (4.12)
g V
7
When the background is warped, the Coulomb potential (4.8) takes the modified form
(5.30) [41], and is extremely flat even at modest separations: see 5.1.
166 4 What is String Inflation?
which does not depend on the background charge distribution8 (y). The
leading term in the scalar potential for a D3-brane is therefore (see 5.1 for
more details)
V (yb ) = 2T3 e4A(yb ) 2T3 1 e4A(yb ) . (4.14)
Computing the trace of the Hessian, we find
2
Mpl
Tr() 2yb e4A(yb ;y) = 2 , (4.15)
T3
where we used (4.13) and (3.47). Thus, the potential for a D3-brane in
the presence of an anti-D3-brane, with no other sources beyond those re-
quired by tadpole cancellation, necessarily has a steep unstable direction,
preventing sustained inflation [41].
Although we have presented the problem in the example of D3-branes,
parallel considerations apply to any scenario in which the backreaction of
a source creates a potential for the motion of some object within the com-
pactification: the instabilities that arise will quickly end inflation. On the
other hand, all realistic models involve additional sources of stress energy
at the very least, to stabilize the moduli and the moduli-stabilizing
contributions can in principle lead to a potential suitable for inflation. This
almost always requires some degree of fine-tuning. To make this fine-tuning
explicit, and thus to obtain a complete inflationary scenario in string the-
ory, rather than a plausibility argument for inflation, requires computing
the moduli potential in extraordinary detail.
8
For discussions of the effects of the background charge, see [516].
9
The actual inflationary instability will generally involve one real component of , e.g. the
real or imaginary part, phase, or magnitude of .
4.2 The Eta Problem 167
10
A criticism of D-term inflation based on consistency conditions in supergravity can be
found in [518].
11
See also [521], in which an assumed Heisenberg symmetry protects the flatness of the
potential.
4.3 Super-Planckian Fields 169
can readily spoil the symmetry (see 2.1.4). Thus, asserting an exact shift
symmetry in supergravity is untenable,12 and the question is how badly the
symmetry is lifted in string theory. In 5.2.1 and 5.4, we will encounter
examples of inflation in string theory that try to exploit shift symmetries to
construct natural models of slow-roll inflation. This is a prime example of
the utility of string theory in assessing ultraviolet-sensitive questions: the
nature of the remnant symmetry can be determined by direct calculation
within string theory. A fair summary is that approximate symmetries are
ubiquitous in string theory, but symmetries that are powerful enough to
resolve the eta problem and make inflation natural are considerably less
common.
12
In certain field theories with special structures, it is possible to suppress all dangerous
symmetry breaking terms to the necessary level. For example, in [241], it was shown
that if the inflaton is the phase of a baryonic operator in SUSY QCD with gauge groups
SU (N 5), symmetry breaking operators only arise at dimension seven or larger. In
this case, the inflaton shift symmetry is an accidental symmetry and symmetry breaking
effects are controlled by gauge symmetry. (The same mechanism controls proton decay
in the Standard Model.) Similarly, coupling the inflaton to a conformal field theory can
suppress the Wilson coefficients of the dangerous operators by RG flow [523].
170 4 What is String Inflation?
In this section we will describe some of the general aspects of the kinematic
and dynamic problems. A definitive treatment of dynamics requires detailed
information about the geometry and potential energy in a metastable com-
pactification, and is therefore deferred to the examples of Chapter 5.
1 Ms2
p1
2 L
< . (4.22)
8 gs `s
It may appear that we can make this field range arbitrarily large by choosing
L `s and/or gs 1. However, what is relevant for the Lyth bound is the
canonical field range in units of the four-dimensional Planck mass (3.47),
1 Ms2 L 6
2
Mpl = . (4.23)
gs2 `s
We find
2
7p
g `s
2 < s . (4.24)
Mpl 8 L
For p < 8, the Planck mass grows faster with L than does, so that in
the limit of theoretical control (L > `s and gs < 1), the field excursion is
sub-Planckian.
The constraint (4.24) is clearly weakest for a high-dimensional brane on
an anisotropic compactification with one large dimension. Consider the
spacetime R1,3 S 1 /Z2 X5 , where X5 is a compact manifold of volume
V5 and the interval S 1 /Z2 has length L. A D8-brane that fills R1,3 and
wraps X5 is then a point particle on S 1 /Z2 . Going through the same logic
as above, one finds
2 gs L
2 < 4 ` , (4.25)
Mpl s
4.3 Super-Planckian Fields 171
13
We thank Juan Maldacena for discussions of this point.
14 0
This is known as a compactification of the type I theory see for example the discus-
sion in [282].
172 4 What is String Inflation?
parametrically increase the field range in a theory where the quantum cor-
rections take the form (4.27). Overcoming this problem requires replacing
the estimate (4.27) with a precise computation in an ultraviolet completion,
and then identifying circumstances in which the scaling differs from (4.27):
see 5.4.1.
15
While the precise change is model-dependent, the ratio of initial to final energies is
2 2 2
generally sizable: for example, Vi /Vf 10 in m chaotic inflation.
4.4 Multi-Field Dynamics 173
4
first relation in (4.28), V MKK , indicates the separation of scales that
could be compatible with V Umod .
While destabilization that leads to runaway decompactification is ruinous,
more controllable backreaction of the inflationary energy on the moduli po-
tential can alter the character of an inflationary model without preventing
sustained inflation. In particular, given sufficiently high barriers around a
local minimum of the moduli potential, a time-dependent inflationary en-
ergy can induce evolution of the moduli within the basin of attraction of
the minimum. Incorporating the motion of the moduli can then change the
form of the inflaton potential, as in the rather general flattening mecha-
nism of [532].16 Thus, although shifts of the moduli do not necessarily end
inflation, their effects must be taken into account.
The twin issues of limited parametric separation and of backreaction by
the inflationary energy are common to all scenarios for large-field inflation in
compactifications of string theory: the problem is simply an outcome of the
high energy scale (1.38), combined with the existence of extra dimensions
with radii greater than the Planck length. Even so, these fundamental
problems take many different guises in explicit constructions, and can be
subtle to identify and extirpate. In Chapter 5, we will encounter these
challenges explicitly: e.g. backreaction by relativistic D-branes in the DBI
model (5.3), and by induced charge on NS5-branes in axion monodromy
models (5.4.2).
16
See also [218, 533], where the slight shift of the overall volume induced by motion of a
D3-brane leads to important corrections to the D3-brane potential.
174 4 What is String Inflation?
17
Curvature invariants are also allowed in principle, but can usually be neglected during
an inflationary phase with H Mpl .
18
Symmetries of the high-scale theory may forbid certain operators, or suppress different
Wilson coefficients to varying degrees, as detailed in 2.1. Incorporating these effects in
the ensemble is straightforward, cf. e.g. [42, 240, 534].
19
See 2.3.2 for an explanation of the cutoff value 6 in small-field inflation.
20
The standard deviation controls the rms size of non-renormalizable contributions to
the potential, and is therefore physical; one can estimate by the general logic of 2.1.
4.4 Multi-Field Dynamics 175
R = f (? , ? ) . (4.30)
In some cases, the entropy perturbations eventually decay and the evolution
reaches an adiabatic limit, where the curvature perturbation can again be
expressed as R = H. After that time, the superhorizon curvature pertur-
bations are conserved. If instead reheating occurs before an adiabatic limit
is reached, the late-time curvature perturbations are extremely sensitive to
the details of reheating, leading to a loss of predictivity.
Single-field slow-roll models automatically predict curvature perturba-
tions that are adiabatic, approximately scale-invariant, and approximately
Gaussian, in excellent agreement with observations. None of these proper-
ties is automatic in a general multi-field model. For m H, the entropy
fluctuations have a strongly scale-dependent spectrum. If these fluctua-
tions give the dominant contribution in (4.30), this can destroy the scale-
invariance of the spectrum of curvature perturbations. Scale-invariance can
be preserved, however, if the couplings of the inflaton to the additional fields
preserve the approximate shift symmetry of the inflaton [239, 490].
Alternative sources for curvature perturbations. Models with multiple light
fields offer alternative mechanisms for generating the observed density per-
176 4 What is String Inflation?
4.5 Reheating
Any complete model of inflation must explain how the energy stored in the
inflaton eventually reaches the visible sector and initiates the hot Big Bang.
There are two basic requirements for the process of reheating: Standard
Model degrees of freedom must be heated to a temperature sufficient for
baryogenesis, and the cosmic history must not be spoiled by overproduction
of relic particles in other sectors. The rich structure of inflationary models
in string theory leads to significant challenges for successful reheating, as
well as a range of novel phenomena, as we now review.21
21
See [544] for a review of reheating in field theory.
4.5 Reheating 177
22
Dark radiation, corresponding to relativistic species in a hidden sector, can have dis-
tinctive signatures: see e.g. [303305, 307].
178 4 What is String Inflation?
23
Most searches for cosmic string lensing involve extragalactic objects (cf. e.g. [574]), but
microlensing of stars within the galaxy [575] (rather than of distant quasars [576]) could
probe very low tensions, particularly if the string loops cluster substantially [577].
24
Cosmic strings of even lower tension might be detectable if they passed through the
Earth [590], causing devastating earthquakes while simultaneously providing a window
on Planck-scale physics.
25
The isotropy of the CMB gives an upper limit on the inflationary scale, and hence
on the tension of cosmic strings that could be produced in a phase transition after
inflation. Moreover, high-tension strings can be excluded by searches for lensing and
for the Kaiser-Stebbins effect.
180 4 What is String Inflation?
coupling gs and the compactification volume. Thus, the string tension can
be low enough to satisfy observational constraints. (A similar argument
applies in the strongly coupled heterotic string [598].) Moreover, in infla-
tionary scenarios involving moving D-branes, reheating typically proceeds
by the condensation of a complex tachyon [599], leading to cosmic string
defects via the Kibble mechanism.
Cosmic superstrings have several important characteristics that distin-
guish them from strings arising as topological defects in perturbative quan-
tum field theories [600602]. In type IIB string theory, there are two ele-
mentary one-dimensional objects: the fundamental string, or F-string, and
the D1-brane, or D-string. These strings can form bound states involving
p F-strings and q D-strings, if p and q are relatively prime [603, 604]. The
resulting (p, q) string has tension [604]
1
q
p,q = (p C0 q)2 + e2 q 2 . (4.31)
20
Networks of (p, q) strings yield scaling solutions [605], just like simpler cos-
mic strings. The intercommutation probabilities of cosmic F-strings and
D-strings can be much smaller than those for field theory cosmic strings,
as carefully examined in [602]. In particular, a colliding pair of strings can
miss each other in the compact dimensions [597, 600, 602], and the string
coupling gs also suppresses the intercommutation probability.
Perhaps the most compelling setting for cosmic superstring production is
warped D-brane inflation [41], in which annihilation of a D3-brane/anti-D3-
brane pair via condensation of a complex tachyon automatically produces
a collection of cosmic strings, and warping provides a natural parametric
mechanism through which the tension can be small enough to obey obser-
vational bounds. The stability and tension of these strings depend on the
the details of the model [601, 606, 607], and we defer further discussion to
5.1.6.
One might hope that cosmic superstrings can be distinguished from strings
arising as topological defects in field theory see [570] for a thorough dis-
cussion of this point. This hope is not entirely unjustified: cosmic super-
strings with P < 1 can be told apart from strings in a perturbative field
theory, which have P 1. Furthermore, the spectrum of tensions (4.31)
appears distinctive. On the other hand, a field theory with SL(2, Z) invari-
ance would reproduce (4.31). More generally, the duality between string
theory and field theory makes it difficult, even in principle, to distinguish
F-strings, D-strings, or (p, q) strings of string theory from corresponding de-
fects in strongly-coupled field theories [570]: for example, the (p, q) strings
produced in warped D-brane inflation can also be viewed as strings of the
dual gauge theory. Even so, the detection of a network of cosmic (p, q)
strings would be an unsurpassed opportunity to probe high-scale physics!
4.6 Inflation in String Theory: a Checklist 181
. For each field with a mass m H, the small mass should be explained
either by fine-tuning of explicitly known, fully specified operators in
the effective theory, or by a symmetry that can be shown to survive
in string theory.
In the next chapter, we will review some of the leading examples of string
inflation. We will see that no model is completely successful on all points
of the above checklist.
5
Examples of String Inflation
183
184 5 Examples of String Inflation
to the curvature of the inflaton potential come from the physical effects that
stabilize the moduli. This was the first of many manifestations of the eta
problem in the context of stabilized string compactifications. The task is
therefore to specify the moduli-stabilizing effects and derive the complete
inflaton potential. We pick up the story at this stage.1
L4 L4
e4A(r) = 1 + , with = 4gs N . (5.2)
r4 (0 )2
This is a simple example of a warped solution, as in (3.36). The solution
has constant dilaton2 and a non-trivial four-form potential
1
This section is based mostly on refs. [41, 42, 217].
2
Recall from 3.1.2 that D3-branes decouple from fluctuations of the dilaton. Moreover,
their backreaction on the metric of an ISD compactification (cf. 3.3.1) is completely
captured by an overall warp factor, as in (5.1). D3-branes are therefore considerably
simpler to treat than branes of other dimensionality.
186 5 Examples of String Inflation
Poincare coordinates,
L2 r2
ds2AdS5 = 2
2 dr + 2 dx dx , (5.4)
r L
we see that (5.1) reduces to AdS5 S 5 for r L.
mobile
D3-brane
Fig. 5.1. Brane inflation in AdS5 . A mobile D3-brane fills four-dimensional space-
time and is pointlike in the extra dimension.
3
We have taken the gauge field strength F2 on the D3-brane worldvolume to vanish,
which corresponds to considering a D3-brane without dissolved D1-brane charge.
5.1 Inflating with Warped Branes 187
For small velocities, r 2 e4A(r) , we can expand the square root to get
1 2
4A()
L () T3 e () , (5.7)
2
where we have defined the canonically-normalized field 2 T3 r2 . From
(5.3), we see that a single D3-brane experiences no force in the anti-de
Sitter background: electrostatic repulsion from the four-form background
exactly cancels the gravitational attraction.
where A {1, 2, 3, 4}. This describes a cone over a base Y5 , which is topolog-
ically but not metrically equivalent to S 2 S 3 . To see this, note that if
z A is a solution to (5.8) then so is z A , with C. Writing z A = xA + iy A ,
the complex equation (5.8) may be recast as three real equations,
1 1
x x = 2 , y y = 2 , xy =0 . (5.9)
2 2
The first equation defines a three-sphere S 3 with radius / 2, while the last
two equations describe a two-sphere S 2 fibered over the S 3 . More precisely,
the base Y5 of the cone is the Einstein manifold5 T 1,1 , which is the coset
space
T 1,1 = [SU (2) SU (2)]/U (1) , (5.10)
4
Mirage cosmology [629] is an alternative to inflation in which the spacetime metric is
the induced metric on a D-brane moving through a background supergravity solution.
Discussions of mirage cosmologies involving D3-branes in warped throat regions include
[630632].
5
An Einstein manifold satisfies Rab gab .
188 5 Examples of String Inflation
2
!2 2
1 1 X 2
d2T 1,1 di + sin2 i d2i ,
X
d + cos i di + (5.11)
9 6
i=1 i=1
where i [0, ], i [0, 2] and [0, 4]. The metric on the conifold
can then be written as
ds2 = k dz dz , (5.13)
4
!2/3
3 X A 2
k(z , z ) = |z | , (5.14)
2
A=1
via k = k.
Deformed conifold.In the singular conifold, the base manifold T 1,1 shrinks
to zero size at zA = 0, and the metric on the cone has a curvature singularity.
To remove the singularity, we consider a small modification of the embedding
condition (5.8),
4
2
= 2 .
X
zA (5.15)
A=1
x x y y = 2 , (5.16)
x x + y y = 2 . (5.17)
where
L4 27
e4A(r) = 1 + with L4 g N (0 )2 . (5.19)
r 4 4 s
For r L, the solution is AdS5 T 1,1 [634].
Warped deformed conifold.Finally, we describe the warped deformed coni-
fold, or Klebanov-Strassler (KS) geometry [426]. This is a noncompact,
smooth solution of type IIB supergravity in which warping is supported by
background fluxes. The KS solution can be obtained by considering the
backreaction of N D3-branes at the tip of the singular conifold, together
with the backreaction of M D5-branes wrapping the collapsed S 2 at the
tip, but we will find it useful to give an alternative presentation in which all
D-branes are replaced by fluxes carrying the associated charges (cf. [635]).
The geometric substrate for the solution is the deformed conifold (5.15),
which contains two independent three-cycles: the S 3 at the tip, known as
the A-cycle, and the Poincare dual three-cycle, known as the B-cycle. The
background three-form fluxes of the KS solution are quantized
Z Z
1 1
F3 = M and H3 = K , (5.20)
(2)2 0 A (2)2 0 B
where M 1 and K 1 are integers. These fluxes give rise to non-trivial
warping. The line element for the KS solution takes the form
where
27
L4 g N (0 )2 , N MK . (5.23)
4 s
Here, rUV is an ultraviolet cutoff, discussed further below. The logarithmic
running of the warp factor corresponds to that seen in the singular warped
conifold solution of [628]. The warp factor eA(r) in (5.21) reaches a minimal
value eA(rIR ) eAIR at the tip, and is given in terms of the flux quanta by
[295]
2K
eAIR = exp . (5.24)
3gs M
The exponential hierarchy is a consequence of the logarithmic running in
(5.22). The KS solution given in (5.21) and (5.22) is the canonical example
of a warped throat geometry, and provides the basis for the most explicit
studies of warped D-brane inflation.
Before proceeding, we should emphasize that the ten-dimensional KS so-
lution, involving a noncompact warped deformed conifold, does not give rise
to dynamical gravity upon dimensional reduction to four dimensions: the
compactification volume, and hence the four-dimensional Planck mass, are
infinite. For model-building purposes, one considers instead a flux compact-
ification containing a finite warped throat region that is well-approximated
by a finite portion of the KS solution, from the tip r = rIR to some ul-
traviolet cutoff r = rUV . Beyond this, the throat attaches to a bulk space,
corresponding to the remainder of the compactification (see fig. 5.2). The
metric of the bulk is poorly characterized in general, but the influence of
the bulk supergravity solution on dynamics in the throat region can be pa-
rameterized very effectively. The validity of the finite throat approximation
was systematically investigated in [42] see 5.1.2.
2
and the volume VB of the bulk space. The Planck mass (3.47), Mpl =
V/gs2 2 , is finite, with V VT + VB . Ignoring the bulk volume gives a lower
bound on the Planck mass,
2
2 N rUV
Mpl > . (5.26)
4 (2 3 )gs (0 )2
5.1 Inflating with Warped Branes 191
4D Perspective
The four-dimensional effective theory can be described by the F-term po-
tential of N = 1 supergravity,
h i
VF = eK K I J DI W DJ W 3|W |2 , (5.34)
where I, J runs over all moduli. We make the standard KKLT assumption
that the complex structure moduli and the dilaton are stabilized at suffi-
ciently high energies. The remaining moduli are then the Kahler moduli Ti
and the brane position moduli z ( = 1, 2, 3). For simplicity of presenta-
tion, we restrict to compactifications with only a single Kahler modulus T ,
but all our considerations generalize to h1,1 I
+ > 1. We define Z {T, z }.
h i h i
K(Z I , Z I ) = 3 ln T + T k(z , z ) 3 ln U (Z I , Z I ) . (5.38)
where in the second equality we have made the dependence on the Planck
mass explicit. We see that the inflaton has a mass of order the Hubble scale,
H 2 V0 /(3Mpl2
). This is how the curvature coupling (5.32) arises in the
effective supergravity description.
bulk CY
D3 D7
warped throat
f (z ) = 0 , (5.42)
f (z ) 1/Nc
A(z ) = A0 . (5.44)
f (0)
15.71 7.85
15.70 7.84
15.69 7.83
15.68 7.82
15.67 7.81
5.47 2.29
5.46 2.28
5.45 2.27
5.44 2.26
5.43 2.25
Fig. 5.3. Example scan through the parameter space of warped D3-brane inflation
(figure adapted from [218]). The scan parameter s is the ratio of the antibrane
energy to the F-term energy before uplifting. Successful inflation occurs in the
gray shaded region.
single explicit embedding is known in which inflation can occur [218, 637].
This is the so-called Kuperstein embedding [639]
f (z1 ) = z1 . (5.45)
In this example, the scalar potential (5.40) receives a correction scaling as
3/2 ( z1 ),
V
VF () V0 + + 3/2 + 02 2 + . (5.46)
Mpl
The last two terms shown in (5.46) contribute to with opposite signs.
Let 0 be the point in field space where the second slow-roll parameter
vanishes, (0 ) = 0. Near this point we have || 1. This is not even
a fine-tuning, but arises dynamically. What does involve fine-tuning is the
requirement that 0 is in the region of control (i.e. inside the warped throat)
and that the potential is monotonic and has a small first derivative (small
) at the same point. If this can be arranged, then we get inflation near
an approximate inflection point. Fig. 5.3 shows an example of a successful
scan in the parameter space of warped D3-brane inflation [218].
10D Perspective
The example above provides an existence proof for inflation in warped throat
geometries, but the setup is too special to provide a good sense for the range
of possibilities. Moreover, the above analysis implicitly assumed that the
physics inside the throat decouples completely from the physics of the bulk,
which as we stressed in 4.2 is rarely the case. Finally, we have modeled the
warped throat region by a finite portion of a noncompact warped Calabi-
Yau cone. This approximation fails where the finite throat is attached to
the remainder of the compactification. In this section, we describe a more
general analysis that addresses these deficiencies.
The essential idea is that all compactification effects i.e. all infor-
mation about moduli stabilization and supersymmetry breaking in the re-
mainder of the compactification can be expressed as non-normalizable
perturbations of the noncompact solution [42, 640],
r
(r) = (rUV ) . (5.47)
rUV
Here, is the deviation of some supergravity field from its value in
the noncompact solution, rUV is the radial location of the ultraviolet end
of the throat, and is the scaling dimension of .6 By determining the
6
In AdS/CFT, this corresponds to the dimension of the operator dual to the perturbation
.
5.1 Inflating with Warped Branes 197
probe D3-brane
+ G + G + , (5.51)
with
G (?6 i)G3 and e4A . (5.52)
At the same time, the three-form flux must satisfy the equation of motion
i d =0.
d + ( + ) (5.53)
2 Im
The solutions to (5.50) can be organized as follows:
where x r/rUV and stands collectively for all five angular coordinates.
We will describe each of the terms in (5.54) in turn.
Constant contributions.The constant V0 represents possible contributions
from distant sources of supersymmetry breaking in the bulk of the com-
pactification, or in other throats that exert negligible forces on the D3-
brane, and only contribute to the inflationary vacuum energy. This situation
corresponds to maximal decoupling of the source of supersymmetry break-
ing from the D3-brane action: the two sectors communicate only through
four-dimensional curvature. As explained in 4.2, complete decoupling of
this sort is very rare. We have in fact made an artificial but convenient
division, using V0 to represent the sum of all8 constant contributions to the
potential, from diverse sources, each of which will in general also contribute
non-constant terms in other categories described below.
7
In comparison to (3.88), we have now allowed the four-dimensional curvature R4 to be
nonvanishing: compare (3.82) and (5.49).
8
In fact, one constant contribution is grouped in VC rather than in V0 : this is the vacuum
energy contributed by the brane-antibrane pair, denoted D0 in (5.55).
5.1 Inflating with Warped Branes 199
Taking into account selection rules [640,641] for the angular quantum num-
bers, the first few scaling dimensions are
3
h = , 2 , 3 , 28 2 , (5.62)
2
where
p
1 (L) 1 + H(j1 , j2 , R + 2) + 4 , (5.64)
p
2 (L) H(j1 , j2 , R) + 4 , (5.65)
p
3 (L) 1 + H(j1 , j2 , R 2) + 4 . (5.66)
5 5
f = 1 , 2 , , 28 , (5.67)
2 2
The total bulk potential (5.57) therefore contains terms with the scaling
dimensions
3 5 7 3
= h , f = 1 , , 28 , 3 , 282 , , 28 , (5.68)
2 2 2 2
A few remarks about the analysis leading to (5.68) are necessary. One
should recognize that (5.59) is non-linear in perturbations of the back-
ground: a linear treatment would capture only the homogeneous solutions
solving (5.58), with dimensions given in (5.62), while the leading term at
small r, corresponding to f = 1 in (5.67), actually arises at quadratic
order in perturbations of three-form flux. This is possible because the per-
turbations corresponding to various supergravity fields do not enter on equal
footing: some perturbations are allowed by the ISD background, and hence
have order-unity perturbations at r = rUV , while other perturbations
are forbidden in the ISD solution, and have perturbations eaT at
r = rUV . These hierarchies can be captured by a careful spurion analysis
[42, 534].
Notice that we again have a contribution scaling as 3/2 , just as in the
four-dimensional analysis. This suggests that the basic phenomenology is
again that of inflection point inflation, and a number of numerical investi-
gations [240, 255, 535, 642, 643] have confirmed this expectation.
5.1 Inflating with Warped Branes 201
9
The physical relevance of the matrix model (5.69) can be understood by comparing it
to the Wigner+Wishart+Wishart model (3.156) of [48]. The positive-definite blocks
AA and B B are consequences of spontaneously broken four-dimensional supersymme-
try: in the limit of unbroken supersymmetry the mass matrix must be positive defi-
nite. The methods used in [42] to construct the ensemble of effective Lagrangians were
inherently ten-dimensional, and made no direct connection to the structure of four-
dimensional N = 1 supersymmetry. The fact that the stability properties enjoined
by four-dimensional supersymmetry nevertheless emerge after the intricate analysis de-
scribed above is encouraging evidence that the entire construction is self-consistent.
202 5 Examples of String Inflation
5.1.4 Reheating
The reheating stage of warped D-brane inflation was carefully examined
in [545, 546, 548, 551, 553], revealing a complex cascade of energy from the
inflaton to the visible sector and to invisible relics. To set the stage, we
remark that a modular approach to reheating is very natural in this con-
text: because the inflaton sector involves a D3-brane in a local geometry,
it is reasonable to identify the warped throat where inflation occurs as one
module, and to situate the Standard Model on D-branes in a different re-
gion of the geometry, either in another warped throat or in the unwarped
bulk region. These model-building choices critically affect the success of
reheating. We will not review all possibilities here, and will emphasize
the interesting two-throat scenario in which the visible sector resides in
a warped throat distinct from that in which inflation occurs. This choice
affords much latitude in model-building, as well as leading to novel phe-
nomenology for reheating. Moreover, if the Standard Model D-branes were
inside the inflationary throat, any relic cosmic strings would quickly disinte-
grate through contact with these D-branes; in the bulk or in another throat,
the D-branes are at a safe distance, and long-lived cosmic strings, with the
associated interesting signatures, are possible.
10
As noted above, some effects of a single light K
ahler modulus were considered in [533],
and the response of the overall volume to the displacement of the D3-brane played a
key role in the stability analysis of [218].
5.1 Inflating with Warped Branes 203
11
See [648] for an analysis of energy transfer in warped reheating via induced motion of
D-branes.
204 5 Examples of String Inflation
annihilation into
closed string loops
end of inflation
3 decay into
2 KK and gravitons
decay into
SM brane modes
4
decay
Fig. 5.5. The stages of reheating after warped D-brane inflation (figure adapted
from [545]).
12
Annihilation can in principle reduce the relic density [551], but only for problematically
small values of the warp factor [553].
5.1 Inflating with Warped Branes 205
5.1.5 Fine-Tuning
Considerable effort has been directed at finding mechanisms that can al-
leviate the fine-tuning of the potential in warped D-brane inflation see
[38,651656]. Here, we will outline a few of the leading approaches. The DBI
mechanism, which turns a steep potential from a liability into an asset, will
be discussed in 5.3. Discrete symmetries can be used to forbid problematic
mass terms [652], and in some cases have been shown to be compatible with
moduli stabilization [657]. Dynamical mechanisms have also been found: it
was shown in [653,655] that if N D3-branes become trapped in a metastable
13
The restriction to irrelevant perturbations in [553] rests on the requirement that the
background throat solution is a good approximation in the infrared. However, [42, 534]
showed that certain relevant perturbations are necessarily present in Klebanov-Strassler
regions of KKLT compactifications. Approximate no-scale symmetry ensures that these
perturbations have exponentially small coefficients and do not destroy the throat. The
effects of such perturbations on Kaluza-Klein relic decays have not been assessed.
206 5 Examples of String Inflation
minimum of the potential in the throat, and sequentially tunnel out, the bar-
rier diminishes with each tunneling event. For favorable parameter values,
the potential for the final D3-brane is an inflationary inflection point.
The second fine-tuning problem of warped D-brane inflation indeed, of
most scenarios for inflation in string theory is that rather special initial
conditions are required for successful inflation to occur. When the poten-
tial is approximately flat in a small fraction of the field space, and is steep
elsewhere, then generic trajectories passing through the would-be inflation-
ary region will overshoot the flat portion without initiating an inflationary
phase, as emphasized long ago in [658]. The DBI kinetic term (see 5.3) has
been argued to ameliorate the overshoot problem [659], though this con-
clusion was challenged by [660]. Negative spatial curvature resulting from
tunneling entirely removes overshooting in certain classes of potential, and
reduces its severity in general [661, 662]. Finally, it was argued in [663] that
the overshooting of an inflection point is ameliorated by particle production
near points in field space where new species become light [664667].
A different perspective on overshooting was given in [240], in which in-
flationary solutions were found by Monte Carlo sampling of the ensemble
of potentials obtained in [42], followed by numerical solution of the six-field
equations of motion. In this setting, the potential was fine tuned by chance,
rather than by hand. Surprisingly, the overshoot problem was absent: for
each inflationary trajectory that was found for a given potential V and for
some fixed initial conditions, an O(1) fraction of the space of possible initial
positions likewise led to prolonged inflation. Thus, while inflation was not
a generic outcome in the joint space of Lagrangians and initial conditions,
for each successful Lagrangian that was found, inflation occurred for generic
initial positions14 of the D3-brane.
5.1.6 Phenomenology
The phenomenology of warped D-brane inflation has been the subject of
intense investigation (e.g. [218, 240, 255, 533, 535, 637, 668670]). In this sec-
tion, we will summarize some of the main conclusions.15 We will start with
the simplified single-field treatment [218, 637] in which the angular degrees
of freedom are integrated out using an adiabatic approximation. This is not
always consistent, as the angular fields can have masses that are smaller
than the inflationary Hubble scale, but serves to develop intuition for the
more complex multi-field dynamics studied in [240, 255, 533, 535, 673, 674].
14
The initial kinetic energy of the D-brane was required to be somewhat smaller than the
initial potential energy.
15
We will emphasize the original scenario [41] in which a D3-brane falls toward the tip of
3
a Klebanov-Strassler throat. Scenarios involving D-branes moving on the S at the tip
of the throat include [671, 672].
5.1 Inflating with Warped Branes 207
We will then present the multi-field results of [240, 255, 535], which incor-
porate the complete potential derived in [42], and follow the full six-field16
dynamics numerically, making no approximation.
Single-Field Expectations
In [218, 637], the six-dimensional field space was analyzed analytically. The
potential was minimized in the angular directions and an effective potential
for the radial direction was determined. As expected, the potential for the
effective radial coordinate has an inflection point. Near the inflection point,
we can write the potential as
" #
1 3
V () V0 1 + 0 + 3 + , (5.71)
Mpl 3! 0 Mpl
where the constants V0 , 0 and 0 can be related to microscopic parameters
of the model [218]. A slow-roll analysis of this potential leads to the following
predictions [218]:
. Power spectrum. The spectral index derived from (5.71) has the
analytic solution [218, 219]
!!
N?2
4 N? 4
ns 1 cot 1+O 2 , (5.72)
Ntot Ntot N? Ntot
where N? corresponds to the number of e-folds between the horizon
exit of the pivot scale and the end of inflation and Ntot denotes the
total number of e-folds, defined as
Z s
1 d 2
Ntot = = . (5.73)
2 Mpl 0 1
The number of e-folds from some initial vev until the end of inflation
at end is
Z !
1 d Ntot ()Ntot
Ne () = = arctan . (5.74)
2 Mpl 2
end
end
For Ntot not much greater than N? 60 the spectrum is strongly blue
and the model is hence ruled out by observations (see fig. 5.6). For
Ntot 2N? , the spectrum on CMB scales is exactly scale-invariant,
while for Ntot > 2N? , the spectrum is red and asymptotes to the lower
limit ns 1 4/N? 0.93 for Ntot 2N? .
16
As explained above, light K
ahler moduli may also evolve during inflation.
208 5 Examples of String Inflation
1.20
1.15
1.10
1.05
200 400 600 800
1.00
50 100
0.95
0.90
Fig. 5.6. Prediction for ns as a function of the total number of e-folds. The gray
band shows the range of ns allowed by Planck.
17
See [675] for another discussion of geometric constraints in warped D-brane inflation,
with implications for eternal inflation.
18
If inflation is driven by the motion of a Dp-brane wrapping a (p3)-cycle, the field range
can be larger than that for a D3-brane [676, 677]. However, arranging for a nearly-flat
potential is challenging, and backreaction of the moving brane can be important.
5.1 Inflating with Warped Branes 209
Multi-Field Effects
As explained in 5.1.3, a proper description of warped D-brane inflation
involves all six D3-brane coordinates (and ultimately any light Kahler mod-
uli). Here, we summarize a few key phenomenological results that emerge
from an intensive Monte Carlo investigation of the dynamics and signa-
tures of the six-field effective theory [240, 255, 535]. A few words about the
methodology are necessary: in [240, 255, 535], scalar potentials were drawn
at random from the ensemble described in 5.1.3, and the equations of mo-
tion were solved numerically beginning from a random initial condition. The
cosmological signatures were then evaluated in the subset of trials that led
to Ne 60 e-folds of inflation.
Fig. 5.7. Multi-field effects on the spectral index in warped D-brane inflation,
versus the mass m of the adiabatic fluctuation (figure adapted from [255]). The
exact tilt, nexact
s , is the result of a six-field numerical calculation making no slow-roll
approximation, while the naive tilt, nnaive s , follows from simply evaluating (2.44)
at horizon exit.
Fig. 5.8. The spectral index in realizations of warped D-brane inflation with sig-
nificant multi-field effects (figure adapted from [255]). The gray band shows the
region allowed at 2 by WMAP7; the Planck constraints are slightly more stringent.
19
Sharp features in the radial profile of the warp factor were argued in [684] to produce
observable signatures in the power spectrum and bispectrum.
212 5 Examples of String Inflation
Cosmic Strings
Cosmic superstrings are one of the most striking signatures of D-brane in-
flation. Following [570], we recall the conditions for cosmic strings to be
cosmologically relevant: the strings must be produced after inflation, re-
main stable over cosmological times, and be observable without already
being excluded. Finally, one may also hope that the strings have distinctive
signatures revealing their origin in string theory. All four conditions can be
met in warped D-brane inflation, as we now explain.
Condensation of the D3-brane/anti-D3-brane tachyon at the end of infla-
tion automatically produces a population of cosmic F-strings and D-strings,
as well as the more general (p, q) string bound states. Whether these strings
are stable depends on whether there are D-branes in the warped throat
where inflation occurs see [601] for a detailed treatment. First of all,
(p, q) strings (including the (1, 0) F-string and (0, 1) D-string) are not BPS
in this setting: the two-forms B and C whose charges the strings carry
are projected out by the orientifold action [601]. Correspondingly, a string
can break apart by coming into contact with its orientifold image. How-
ever, in the generic situation in which there are no orientifold fixed planes
within the throat itself, a string has to fluctuate out of the throat to meet
its image in the image throat. This is an exponentially slow process, as the
potential due to the warp factor confines the strings to the bottom of their
respective throats, and for practical purposes breakage via the orientifold
image can be ignored [601]. A more significant risk comes from D3-branes or
anti-D3-branes in the inflationary throat, which could serve as the substrate
for the Standard Model [685] or as a source of supersymmetry-breaking en-
ergy [425]. If any D3-branes or anti-D3-branes are present, cosmic strings
fragment immediately and are cosmologically irrelevant. If D7-branes are
present but D3-branes and anti-D3-branes are not, the D-string remains
stable [601].
The spectrum of tensions of (p, q) strings in a warped throat was obtained
in [686]:
s
e2AIR q 2 bM 2 2 (p qC0 )
T(p,q) + sin , (5.78)
20 gs2 M
where eAIR is the warp factor at the tip of the throat, M is the flux on
the A-cycle, and b 0.93 is a constant arising in the Klebanov-Strassler
solution. This result is primarily governed by the warp factor, which can
be exponentially small. Hence, if the warp factor were a free parameter,
it would be easy to ensure that the cosmic string tension is low enough to
satisfy any conceivable observational bound. However, the warp factor in
the inflationary throat determines the scale of the inflaton potential, and is
5.2 Inflating with Unwarped Branes 213
SUSY-breaking
D7-brane
%"+,#&'02 volume-stabilizing
mobile D3-brane
D7-branes
20
See [689691] for analyses of consistency conditions for Fayet-Iliopoulos terms in super-
gravity, and [692] for the implications for cosmic strings.
21
An important alternative means of breaking supersymmetry is the addition of an anti-
D3-brane in a warped region see [687] for a comprehensive discussion of the D3/D7
model with antibrane supersymmetry breaking.
5.2 Inflating with Unwarped Branes 215
22
A very different perspective on this fact was recently given in [693].
216 5 Examples of String Inflation
tive action, the D3-brane potential takes the form of a nearly flat trough
oriented along the symmetry direction. However, the nonperturbative su-
perpotential, which is critical in the stabilization of the volume, necessarily
depends on the D3-brane position, contrary to our assumption above. The
one-loop correction to the gauge kinetic function for the D7-brane gauge
theory was computed explicitly in [336], and was found to depend on the
D3-brane position, so that the gaugino condensate superpotential likewise
depends on the D3-brane location. For D7-branes with gauge group SU (Nc )
at position zD7 = in T 2 /Z2 , one finds [336]
h i1/Nc 2T /Nc
W = W0 + 1 2(z1 + ), 1 2(z1 ), e , (5.83)
23
A T-dual configuration, in which the inflationary coordinate is a Wilson line, has been
investigated in [698] (see also [699]).
218 5 Examples of String Inflation
Fig. 5.10. In fluxbrane inflation [614, 615], the inflaton coordinate is the effective
separation of a pair of intersecting D7-branes (figure adapted from [614]).
24
This setup is T-dual to a configuration of branes at angles: to see this, take the com-
pactification to be a torus and T-dualize along a circle in the two-cycle threaded by the
flux F.
25
In an alternative parameter regime for fluxbrane inflation, the dominant non-constant
term in the potential is sinusoidal [251], instead of logarithmic as in (5.87).
5.2 Inflating with Unwarped Branes 219
26
D3-branes, in contrast, do enjoy a moduli space in the leading order no-scale com-
pactifications of [295], but at this same order the Kahler moduli are unstabilized. The
challenge described in 5.1 is that the nonperturbative effects that lift the K
ahler moduli
inevitably spoil the flatness of the D3-brane potential.
220 5 Examples of String Inflation
27
It was argued in [701] that the tensor-to-scalar ratio r can be large in multi-M5-brane
inflation.
28
Vector bundle moduli are not necessarily stabilized, but are sometimes assumed to be
absent.
29
The difficulties inherent in constructing parametrically controlled heterotic vacua with
H3 were appreciated many years ago [331, 332], and have been only partially overcome:
see [703, 704].
5.2 Inflating with Unwarped Branes 221
5.2.4 Phenomenology
Once moduli stabilization is properly incorporated, inflation in the D3/D7
model remains possible, but necessarily involves fine-tuning. Equipped with
the global form of the nonperturbative superpotential thanks to the world-
sheet calculation of [336], the authors of [613, 687] systematically analyzed
the potential in search of inflationary regions. Two qualitatively different
scenarios were found:
. Saddle-point inflation.If the condensate responsible for Kahler mod-
uli stabilization is assumed to arise exclusively on a stack of D7-branes
near a single fixed point of T 2 /Z2 , then after fine-tuning of the pa-
rameters, the potential for a D3-brane at an approximately antipo-
dal location in the torus can develop an unstable saddle point. (For
this scenario, it is essential that the primary source of supersymmetry
breaking is an anti-D3-brane in a warped region.) The resulting model
has r 1 and ns . 0.95. The characteristic redness of the spectrum
in saddle-point models of this form is discussed in [705].
. Inflection point inflation.If the dominant force on the D3-brane
comes from interactions with supersymmetry-breaking fluxes on a D7-
brane, as in [609], then the potential can be fine-tuned to have an
inflationary inflection point, with phenomenology broadly similar to
that described in 5.1.6. The potential takes the form (5.85), incor-
porating a Coleman-Weinberg term, as well as quadratic and quartic
terms from moduli stabilization. When the quartic terms are signifi-
cant, the fine-tuning for inflation is extreme, and was argued in [687]
to be at the level of one part in 106 . On the other hand, [613] exhibit
parameter ranges in which the moduli contribution is approximately
quadratic and the fine-tuning is milder.
The kinematical field range of the canonically-normalized inflaton in
D3/D7 inflation can be super-Planckian, > Mpl , if the T 2 /Z2 is highly
anisotropic [613]. For a rectangular torus with side lengths L1 and L2 , we
have p
Mpl Vol(K3)L1 L2 , (5.89)
while the field range along the side of length L1 has the parametric depen-
dence p
1 L1 /L2 . (5.90)
In [613], it was argued that one can take L1 /L2 to be large enough so
that 1 > Mpl , without compromising computability. This fact is con-
sistent with the general arguments made in 4.3, where anisotropy of the
compactification was the only plausible route to a parametrically controlled
super-Planckian field range for a D-brane. Even so, this observation has not
222 5 Examples of String Inflation
30
This section is based mostly on [38, 40].
5.3 Inflating with Relativistic Branes 223
31
The constant is proportional, but not equal, to the standard t Hooft coupling in
5 2
AdS5 S , t gYM N = 4gs N .
32 1,1
Such a region actually corresponds to a section of AdS5 T , up to logarithmic
corrections, but the angular manifold is immaterial at present.
224 5 Examples of String Inflation
4A(r) L4 4 4 gs
e with L 4
N (0 )2 , (5.96)
r4 Vol(X5 )
where Vol(X5 ) denotes the volume of X5 (in string units). For a throat of
the form (5.95), the range of the canonically-normalized D3-brane position
is given, as in (5.28), by [243]
2
, (5.97)
Mpl N
and in particular is independent of X5 . We will see in 5.3.5 that if one
manages to achieve a DBI phase, (5.97) provides a stringent upper bound
on the tensor-to-scalar ratio.
The UV model.We will refer to the UV model as the situation in which
a D3-brane moves into the warped region, i.e. toward small , from the
ultraviolet [38, 40].
The IR model.An interesting alternative is the IR model, in which in-
flation occurs as the D3-brane leaves the tip region and moves toward the
ultraviolet end of the throat [711713]. The initial conditions for the IR
model are very appealing [711]. Suppose that p anti-D3-branes are intro-
duced into a KS throat region. If p . 0.08M , with M the flux quantum
number defined in (5.20), the anti-D3-branes form a metastable configura-
tion at the tip [425]. Over an exponentially long timescale, this state can
decay: the anti-D3-branes annihilate against flux, liberating (M p) D3-
branes. The observation of [711] is that the D3-brane potential arising from
33
For a generalization of DBI inflation to arbitrary warp factor, assuming an appropriate
potential, see [709, 710].
5.3 Inflating with Relativistic Branes 225
moduli stabilization may drive some or all of these D3-branes to move out
of the throat region, and during this process a phase of DBI inflation can
occur. A simple model of the radial potential is
1
V () = V0 H 2 2 , (5.98)
2
where V0 is a constant, and in generic configurations,34 1 . . 1. The IR
model corresponds to > 0. For a discussion of obstacles to a computable
realization of the IR model in string theory, see 5.3.4.
Relativistic dynamics.For a spatially-homogeneous D3-brane, i.e. for =
(t), it is natural to define a Lorentz factor, by analogy to relativistic
particle dynamics:
!1/2
2
1 . (5.99)
T ()
The requirement that be real enforces a speed limit on the motion of the
probe D3-brane:
2 < T () . (5.100)
Notice that the bound is independent of the properties of the potential and
becomes stronger in regions of strong warping, where e4A() 1 and hence
2 T3 .
In 5.1, we studied non-relativistic D3-brane motion, corresponding to
1: expanding the square root in (5.6) then led to the two-derivative
action (5.7). DBI inflation operates in the regime of relativistic brane dy-
namics, with 1, where higher-derivative terms in (5.92) cannot be
neglected. Varying (5.92) with respect to the four-dimensional metric gives
the stress-energy tensor sourced by the D3-brane. This corresponds to the
stress-energy of a perfect fluid, with energy density and pressure given by
= 1 T +V , (5.101)
T
P = 1 V . (5.102)
Coupling to gravity gives the Friedmann equation
2
H 2 = 1 T + V () ,
3Mpl (5.103)
and the continuity equation
2 0
2Mpl H
= , (5.104)
34
Potentials that tend to expel a D3-brane from the infrared are quite common in the
ensemble obtained in [42], but are far more complicated than (5.98), involving all five
angular directions and an array of competing terms (see [240]).
226 5 Examples of String Inflation
Notice the factors of 1 in both and . For large , the slow-roll pa-
rameters are therefore suppressed relative to the expectation derived from
the non-relativistic limit. This leads to the intriguing possibility of achiev-
ing inflation even for potentials that naively seem to be too steep to drive
prolonged inflation.
Accelerated expansion occurs if the potential energy dominates over the
kinetic energy. Demanding that V () is the leading term on the right-hand
side of (5.101) gives the condition
V
1. (5.108)
T
Thus, DBI inflation can occur near the location only if the potential is
large in local string units. Next, a defining requirement for a DBI phase is
that the D3-brane is relativistic.35 Using (5.108) in (5.103) and (5.105), we
find
2 V
2 = 1 , (5.109)
3 T
where 21 Mpl2
(V 0 /V )2 . We observe that although (5.109) involves and
can plausibly be satisfied by making the potential very steep, the condition
(5.108) is independent of the functional form of the potential. We will see
below that (5.108) presents a serious obstacle in the search for a consistent
embedding in string theory.
Analytical [38, 40] and numerical [668, 708] studies have shown that for
suitable potentials V (), the DBI Lagrangian (5.92) can indeed support
an inflationary phase in which the non-trivial kinetic term plays a crucial
role. However, before describing the phenomenology of DBI inflation, one
should first ask whether the DBI Lagrangian (5.92) gives an accurate and
consistent representation of the physics of a relativistic D3-brane. There are
35 1
In the remainder of this section, we will work to leading order in 1.
5.3 Inflating with Relativistic Branes 227
36
This argument was first made in [504] and further elaborated in [714, 715].
228 5 Examples of String Inflation
which constrain the form of the action. First, we note that the unbroken
four-dimensional Lorentz symmetry and the nonlinearly realized dilatation
symmetry (5.110) imply
Z
S = d4 x 4 f ()2 /4 + , (5.112)
37
Here, we assume that supersymmetry is only broken spontaneously during inflation, so
it still constrains the form of the action.
5.3 Inflating with Relativistic Branes 229
38
The terminology is arguably more appropriate to N = 2 theories, where the Coulomb
branch is distinguished from the Higgs branch because the former is parameterized by
scalars in vector multiplets and the latter is parameterized by scalars in hypermultiplets.
The important point is just that motion on the Coulomb branch does not change the
rank of the gauge group, but can change the rank of the non-Abelian part of the gauge
group, i.e. U (N ) U (N 1) U (1), while motion on the Higgs branch can change the
total rank.
230 5 Examples of String Inflation
39
Identifying an alternative setting for DBI inflation in a string compactification would
be very interesting, and would undoubtedly lead to modified microphysical constraints,
but we are not aware of any complete example.
5.3 Inflating with Relativistic Branes 231
in realizations consistent with the field range bound (5.97) of [243], and
with observational constraints on ns , the Lorentz factor was bounded
by 1 < 107 .
A remark about the Coulomb potential is relevant here. The poten-
tial VC () given in (5.30) applies to a KS throat (in the AdS5 T 1,1
approximation). For a cone over X5 , one finds instead
D0
VC () = D0 1 . (5.114)
4Vol(X5 ) 4
V () m2
= . (5.115)
T () 2 2
Consider the effect of hidden sector supersymmetry breaking on the
Kaluza-Klein spectrum of the throat. Barring an efficient sequestering
mechanism, the Kaluza-Klein modes will acquire masses MKK m.
Since the lightest Kaluza-Klein modes in an undistorted throat have
1 AIR r
MKK e IR2 , (5.116)
L L
supersymmetry breaking will typically cut off the throat at
rIR mL2 . (5.117)
232 5 Examples of String Inflation
EFT N 1/8
Ne , (5.122)
where as usual N is the D3-brane charge of the throat and was
defined in (5.98). Unless NeEFT 60, the observed CMB perturba-
tions will be dictated by uncomputable fluctuations of massive strings.
While this is a fascinating possibility that escapes the confines of the
effective field theory of a finite number of fields, no meaningful pre-
dictions are possible at present. For generic potentials with 1,
solving the horizon problem without encountering uncomputable per-
turbations requires N & 1014 ; moreover, because Ne 1 (with Ne
the number of e-folds) is required for a consistent DBI phase [720], we
must have N & 107 regardless of the value of . D3-brane charges of
this magnitude are difficult to realize in compact spaces.
5.3.5 Phenomenology
Despite the many apparent obstacles to realizing DBI inflation in a consis-
tent string compactification, the DBI scenario is a leading example of a field-
theoretic inflationary mechanism that is underpinned by the symmetries of
an ultraviolet theory. Many authors have deferred the question of explicit
ultraviolet completion, and directly investigated the rich phenomenology
40
While alternative means of shedding energy could give rise to interesting cosmologi-
cal scenarios, the dynamics would not be governed by (5.92) alone, and much further
analysis would be required. See 5.6 for several examples of dissipative models.
234 5 Examples of String Inflation
cs P,X = 1 . (5.129)
A = 1 . (5.131)
equil 1 c2s
fNL = 0.27 + 0.08A , (5.132)
c2s
ortho 1 c2s
fNL = + 0.02 + 0.02A . (5.133)
c2s
For the DBI Lagrangian, the orthogonal component is small and the
amplitude of equilateral non-Gaussianity becomes
equil 35 1 35 2
fNL = 2 1 , (5.134)
108 cs 108
41 7
The bound r < 10 /Vol(X5 ) can be derived by combining (5.28) with the assumption
equil
that fNL & 1. In other words, DBI inflation in a Calabi-Yau cone cannot have both
detectable non-Gaussianity and detectable tensors [722].
42
This result is independent of the potential and the warp factor, unlike other observables
such as ns .
236 5 Examples of String Inflation
43
For an analysis of DBI inflation with N D3-branes, see [723].
5.4 Inflating with Axions 237
44
The transfer from entropy perturbations to curvature perturbations can depend on
the physics of reheating. Investigations of the distinctive features of reheating in DBI
inflation include [559, 731, 732].
45
The constraint on the decay constant depends on the choice of prior. The result we
have cited here is for a uniform prior on log(f ) [9].
238 5 Examples of String Inflation
Two axions
Consider two axion fields 1 and 2 with decay constants f1 and f2 , re-
spectively. Suppose that these axions couple to linear combinations of two
confining non-Abelian gauge groups a and b, with the following Lagrangian
density [734]
2 i
X i cia h
(a) (a)
i cib h
(b) (b)
L Tr F F + Tr F F , (5.143)
f 32 2 32 2
i=1 i
where the coefficients cia = {c1a , c2a } and cib = {c1b , c2b } are dimensionless.
In terms of the dynamical scales a and b of the two gauge groups, the
potential for the axions is
4 1 2 4 1 2
V = a 1 cos c1a + c2a + b 1 cos c1b + c2b .
f1 f2 f1 f2
(5.144)
The central observation of [734] is that if
c1a c
= 1b , (5.145)
c2a c2b
then one linear combination of the axions is unlifted, and effectively has
infinite range.46 When (5.145) is approximately satisfied, has a decay
constant f that can be much larger than f1 and f2 . In particular, for
sufficiently precise alignment, one can have f > Mpl with f1 , f2 Mpl .
For the simple case where c1a = c1b = 1 and 4a 4b , one finds
1/2
c22a f12 + f22 f c f
f =
c2b c2a
, with = 2 22 2 2a 12 1 . (5.146)
c f +f 2a 1 2
46
A related mechanism is used in racetrack constructionssee 5.5.1.
5.4 Inflating with Axions 239
effects in the gauge groups given in (5.143). Whether other effects lift the
flat direction is an important question: in particular, it would be valuable
to construct an explicit example in a stabilized string compactification and
ascertain whether any nonperturbative effects associated with moduli stabi-
lization, which might be unrelated to the confining gauge groups in (5.143),
spoil the flatness of the inflaton direction. (For a closely related idea, see
[254], as discussed in 5.4.2 below.)
Compared to the single axion case, each individual axion feels enhanced
Hubble friction, suggesting that one might be able to achieve a friction-
dominated situation even without any axion being at a super-Planckian
distance from the minimum.
Let us begin by considering the simple case in which all the axions have
equal masses mi = 2i /fi m. Moreover, let us consider small displace-
ments from the minimum, i fi . The collective excitation 2 i 2i
P
then has the potential
1
V () = m2 2 . (5.149)
2
47
This section is based on [483, 530, 735].
48
A different approach to assisted inflation in string theory is M-flation [736, 737].
240 5 Examples of String Inflation
Successful inflation requires > Mpl , but this does not mean that the
individual axion vevs i have to be large: for sufficiently
large N , the indi-
vidual displacements can be sub-Planckian, i / N < Mpl . In typical
examples the required number of axions is N & O(103 ) [530].
One might object that quantum gravity constraints should restrict the
range of to be sub-Planckian, just like the ranges of the i , because
in moving from individual displacements i to the collective field , one
has merely changed from Cartesian to spherical polar coordinates, which
should be immaterial unless some physical effect is sensitive to the change
of coordinates. This concern is unfounded: the N discrete axionic shift
symmetries i 7 i + 2fi that persist at the nonperturbative level do in
fact define a preferred coordinate system, the Cartesian one. (After the
periodic identifications, the axion field space is an N -torus, with individual
radii fi < Mpl .) In string theory constructions (see below), only the fi are
directly constrained.
A much graver concern is that loops of the N light axion fields renormalize
the Planck mass: on general grounds one expects [530]49
2 N
Mpl 2UV , (5.150)
16 2
where UV is the ultraviolet
cutoff. Because the collective field displace-
ment scales as N , with the mean of the individual displace-
ments,
while the correction to the Planck mass in (5.150) also scales as
N , we conclude that one cannot obtain a parametrically super-Planckian
displacement purely by working at large N . Instead, one must grapple with
the ultraviolet-sensitive details, e.g. by refining the field-theoretic estimate
(5.150) through a computation in string theory. To learn more, we now turn
to a string theory realization of N-flation [483, 530, 735].
N-flation in type IIB string theory.As an explicit realization of N-flation in
1,1
string theory, we consider a KKLT compactification in which h+ N 1
complexified K ahler moduli Ti are stabilized by nonperturbative effects. As
explained in 3.2, the associated axions i , i = 1, . . . , N , correspond to the
integrals of C4 over orientifold-even four-cycles, cf. eq. (3.69). For simplicity
of presentation, we take h1,1
= 0, so that by (3.67), Ti = i + ii , with i a
50
real four-cycle volume.
49
A counterpoint to this finding appears in [738], where it is argued that the portion of
the eta problem arising from loops of light fields is actually suppressed at large N .
50
Another interesting and rather explicit construction with similar qualitative features
1,1
works with h N 1, taking the inflationary axions to arise from the dimensional
reduction of the R-R two-form potential, cf. eq. (3.62) [735]. One advantage of the
model of [735] is that it is comparatively straightforward to arrange that the K ahler
moduli masses are larger than the inflaton mass.
5.4 Inflating with Axions 241
51
At the energy scales of interest, the dilaton and the complex structure moduli are
already stabilized, and W0 and Ai are constants.
242 5 Examples of String Inflation
a simple expression for the axion mass spectrum [483]. The mass matrix
belongs to the Wishart ensemble described in 3.5.3, and the mass spectrum
is given by the Marcenko-Pastur law (3.153),
1
q
(m2 ) = 2 2 (+ m2 )(m2 ) . (5.155)
2N m
Here, N 2 (1 P/N )2 , where, as above, N = h1,1 2,1
p
+ , while P = h is
the number of complex structure moduli, and controls the typical scale of
terms in the moduli potentialsee [483] for a detailed explanation.52 The
eigenvalue spectrum (5.155) is shown in fig. 5.11.
0 1 2 3 4
Fig. 5.11. The eigenvalue spectrum of a Wishart matrix, given by the Marcenko-
Pastur law (5.155) (figure adapted from [48]). The spectrum of axion masses around
a KKLT minimum is well-described by this law when N h1,1 + 1. The curve
is the analytic result, while the histogram is the result of simulations, both for
N = P = 200.
52
One further assumption implicit in obtaining (5.155) in a KKLT compactification is
that the gravitino mass m3/2 is small compared to the scale of supersymmetric masses:
see [48, 476] for discussions of how this could be achieved.
5.4 Inflating with Axions 243
53
A potential confusion is that the leading correction in (5.156) arises as a loop effect on
the worldsheet, not from loops of light moduli in spacetime, but nevertheless involves
the number of moduli. As such, it corresponds to a non-renormalizable term in the
effective theory, as in 2.3.2, rather than a radiative correction as in (5.150), cf. 2.3.1.
However, corrections at higher order in gs may be expected to involve actual loops of
the light fields, in closer analogy to (5.150).
244 5 Examples of String Inflation
54
One cannot work at arbitrarily weak coupling and large volume, because in this limit
the individual decay constants fi are parametrically small compared to Mpl : see 3.2.3.
55
This section is based on [34, 751].
56
The first example of monodromy inflation constructed in string theory involves a D4-
brane in a nilmanifold compactification of type IIA string theory [751], and relies on a
scenario for moduli stabilization that is rather different from that presented in 3.3.3.
We will discuss the scenario of [751] in 5.6.1.
5.4 Inflating with Axions 245
Performing the integral over the two-cycle, we obtain the potential for the
axion in the four-dimensional effective theory:
%
q
V (b) = 6 0 2 (2)2 `4 + b2 , (5.159)
(2) gs ( )
where ` is the size of 2 in string units, and the dimensionless number %
encodes a possible dependence on the warp factor. The presence of the
brane has broken the axion shift symmetry, b 7 b + (2)2 . We say that the
brane has generated a monodromy for the axion. For large values of the
axion vev, b `2 , the potential is linear, V (b) b.
A similar effect occurs if the D5-brane is replaced by an NS5-brane.
R The
wrapped NS5-brane now produces a monodromy for the axion c 10 C2 .
2
Dimensional reduction of the action for the NS5-brane introduces the fol-
lowing potential for the c axion
%
q
V (c) = 6 2 0 2 (2)2 `4 + gs2 c2 . (5.160)
(2) gs ( )
Axion monodromy inflation.In both cases, inflation can occur if the ax-
ion (b or c) has a large initial vev. The Lagrangian for the canonically-
normalized field is
1 1 %
L = ()2 3 , with 3 , (5.161)
2 f (2) gs (0 )2
6
57
For related field-theoretic constructions, see [753, 754].
5.4 Inflating with Axions 247
58
That is, the D5-brane and anti-D5-brane wrap distinct, well-separated representatives
of the same homology class.
248 5 Examples of String Inflation
NS5-brane
anti-NS5-brane
Fig. 5.12. The integral of the two-form C2 over a two-cycle 2 defines the c axion.
In the presence of a wrapped NS5-brane this develops a monodromy. An anti-NS5-
brane is required by Gausss law on the compact space. The entire configuration
should be situated in a warped region, and have a distant orientifold image (not
shown). In the lower figure, the two-cycles are represented by circles.
59
The precise condition on the triple intersection form is ci = 0 i, where denotes
the orientifold-odd two-form corresponding to the inflaton, i.e. c = G (no sum), and
i indexes all four-cycles stabilized by Euclidean D3-branes.
60
This is easily understood if one recalls that the backreaction of D3-branes in flat space
5
leads to the AdS5 S geometry.
61
The corresponding calculation for a D3-brane was performed in [335] see also the
discussion following eq. (4.10).
252 5 Examples of String Inflation
where the first term comes from the DBI action and Vb stands for the
moduli potential. The K ahler moduli Ti depend on c via the warp
factor,
Z
Re(Ti ) = d4 y g e4A(y;c) . (5.174)
i
4
In generic configurations, the resulting c-dependence of the moduli
potential is large enough to invalidate the derivation made in the probe
approximation.
Fortunately, there is a mechanism that provides parametric suppres-
sion of this problematic backreaction effect. In order to satisfy Gausss
law, it was necessary to introduce an anti-NS5-brane, in addition to
the NS5-brane: the induced charges on the brane and antibrane are
equal and opposite. Thus, if the NS5-brane and anti-NS5-brane are
relatively near to each other compared to their distance from the rel-
evant four-cycles, the net flux of F5 past the four-cycle will be sup-
pressed, and the correction to the warp factor will be correspondingly
small. Instead of seeing a monopole D3-brane charge, the four-cycles
see only a dipole. A concrete realization of this protective mechanism
involves placing an NS5-brane and an anti-NS5-brane in a common
warped throat, as discussed in detail in [36].62
To recap, backreaction of the induced D3-brane charge that is ulti-
mately responsible for the inflationary energy leads to a correction
to the warp factor that affects the scalar potential by modifying the
Euclidean D3-brane action. This is consistent with the general argu-
ments, in that the breaking of the axionic shift symmetry by back-
reaction effects is, strictly speaking, nonperturbatively small, being
proportional to exp(Tp Vol(p )). However, it is essential to under-
stand that the moduli potential, and the inflationary vacuum energy,
are necessarily nonperturbatively small in the same sense. For the
backreaction to be a small correction, the geometry must be arranged
to respect an additional approximate symmetry, e.g. by situating the
fivebrane pair at the bottom of a warped throat, as noted above. The
original axion shift symmetry, on its own, does not suffice to guarantee
a flat potential.
. Backreaction from the NS5-branes.There is another backreaction
effect that presents a possible concern [757]. The NS5-brane/anti-
NS5-brane pair explicitly breaks supersymmetry, and moreover either
member of the pair, in isolation, sources a fivebrane tadpole that is
62
A precise computation of the axion decay constant in such a setting is an open problem.
If the decay constant is significantly reduced by warping, the requisite induced charge
increases, intensifying the problem of backreaction.
5.4 Inflating with Axions 253
Reheating
Reheating in a string theory model with a shift-symmetric inflaton poses
particular difficulties, as explored in [557, 560, 562, 564]. The general issue
is that the shift symmetry that protects the inflaton simultaneously limits
the couplings of the inflaton to the visible sector. In a naive model contain-
ing only the inflaton, the visible sector, and general relativity with no
additional sectors associated with the ultraviolet completion of gravity
the primary consequence would be slow reheating, which is not necessarily
fatal. However, if the inflaton couples at least as strongly to hidden sector
fields as it does to the visible sector, which is frequently the case in models
of closed string inflation, then problematic reheating of the hidden sectors
is difficult to avoid.
At present, no concrete results concerning reheating are available for the
254 5 Examples of String Inflation
5.4.3 Phenomenology
The phenomenology of axion inflation can be extremely rich.63 Besides the
model-independent gravitational wave signal, a host of additional model-
dependent signatures have been explored, including oscillations in the power
spectrum [36], deviations from scale-invariance [759], non-Gaussianity [37,
760], chiral gravitational waves [260, 759], and primordial black holes [761].
Since all of these effects are tied to the underlying axion shift symmetry, one
has the hope of finding correlated signatures across different observational
channels.
Signatures of nonperturbative effects.At leading order in the instanton
expansion, the Lagrangian for axion monodromy inflation takes the form
1 2 4
L = () V0 () cos . (5.175)
2 f
where V0 () 4p p . Since the scale is generated nonperturbatively,
the modulations can quite naturally be exponentially small, but it is also
possible that the modulations could be large enough to spoil the mono-
tonicity of the inflaton potential. Here, we assume that happens to be
large enough to be phenomenologically interesting, but not large enough to
dominate the evolution. The monotonicity constraint is
4
b? <1. (5.176)
V00 (? )f
This parameter depends on the inflaton vev, unless the potential is linear.
Here, we have evaluated it at = ? , the value of the inflaton when the
pivot scale k = k? exits the horizon.
Before proceeding, we point out that f /Mpl is bounded from below, for
two reasons. First, requiring that the 0 expansion is under good control
(in the NS5-brane construction [34] with p = 1) implies [36]
f2 gs
2 > . (5.177)
Mpl (2)3 V
Another bound on f arises from requiring that the effective theory of the
fluctuations is weakly coupled for the parameter values of interest [762].
63
See [758] for a recent review.
5.4 Inflating with Axions 255
Control over the observational predictions requires that the oscillation fre-
2 1/2
quency = (2Mpl H) /f is much smaller than the unitarity bound 4f .
This implies
H
f (2R )1/2 . (5.178)
2
Using (1.37), we get
r
f r 1/2 r 1/2
R 4.5 104 . (5.179)
Mpl 16 0.07
where K k1 + k2 + k3 and
3/2
res 3 2 2?
fNL b? . (5.184)
8 f
The ellipses in (5.183) stand for terms that are suppressed by higher
powers of the slow-roll parameters or bypositive powers of f / 2? .
Observable non-Gaussianity requires f / 2? 1. In this case, the
second term in (5.183) is suppressed relative to the first term, except
in the squeezed limit (where it ensures that Maldacenas consistency
relation (1.70) holds). The unitarity bound (5.179) implies an upper
res
bound on fNL ,
3
res
fNL (2R )3/4 3 103 . (5.185)
2
Because the bispectrum is oscillating, it is nearly orthogonal to the
standard bispectrum templates. It is therefore barely constrained by
the present bispectrum results (1.72)(1.74), and a dedicated analysis
res
is required to put meaningful constraints on the parameter fNL .
64
Using WMAP9, it was found that a modulation with log10 (f /Mpl ) = 3.38 improves the
2
fit by = 19 [764], but the frequency of this signal coincides with the unitarity bound
(5.179). Moreover, the signal does not seem to be present in the Planck data [765, 766].
65
We will assume 1. The case 1 is discussed in [767], while estimates for in
type IIB string theory appear in [759]. Both bottom-up and top-down naturalness of
the regime 1 remain to be established.
5.4 Inflating with Axions 257
where F is the gauge field strength. This coupling respects the shift sym-
metry of the inflaton, as for constant the operator is a total derivative.
When the inflaton has a time-dependent vev, (t), the conformal invariance
of the gauge field is broken. This leads to production of gauge field quanta
during inflation. To see this, consider the equation of motion for the two
polarization modes of the gauge field (in Coulomb gauge):
!
2 2
2 + k 2aHk A (, k) = 0 , where
2f H
. (5.187)
We see that one of the helicities of the gauge field experiences tachyonic
growth for k/(aH) < 2. For > 0, the unstable mode is A+ . Most of the
power in the produced gauge field is in modes with (8)1 < k/(aH) < 2.
In this regime, the solution can be written as [767]
1
k
1/4
A+ (, k) ' e2 2k/(aH) . (5.188)
2k 2aH
The coupling of this solution to the inflaton, via (5.186), leads to a number
of observational signatures:
. Equilateral non-Gaussianity.The gauge field non-linearities in the
operator (5.186) source non-Gaussian inflaton fluctuation (this may
be thought of as an inverse decay, A + A ). The bispectrum is
of equilateral type and has the amplitude [760]
equil 6R,0
fNL ' f3 ()e6 , (5.189)
4R
where R,0 stands for R |=0 = H 2 /(2 ).
The function f3 () is
determined numerically, but has the following limits:
f3 () = 2.8 107 9 for 1 , (5.190)
8 8.1
f3 () 7.4 10 for 2<<3. (5.191)
equil
We see that fNL is exponentially sensitive to the model parameter
. Let us denote by ? the value of at the pivot scale k? = 0.002
Mpc1 . Using the seven-year WMAP data, ref. [768] found ? < 2.45
(95% CL). In terms of the axion decay constant, this corresponds to
H
f> . (5.192)
10 R
Using (1.37), this can be written as
r
f r r 1/2
> 2 102 . (5.193)
Mpl 10 2 1 0.07
258 5 Examples of String Inflation
where
66
This section is based mostly on [43, 357, 620]. For a recent review see [30].
67
For an explanation of the exponent 2 in (5.199), which differs from the result given in
[356], see [41].
68
Racetrack-type superpotentials for K ahler moduli have also been argued to yield inflec-
tion point inflation [773].
260 5 Examples of String Inflation
69
The example given in [618] has N = 90 and M = 100, as well as fine-tuned values for
W0 , A, B, %. In the more explicit construction of [619] (building on moduli stabilization
results of [358]), successful inflation was found for N = 40 and M = 258, again with
fine-tuning of the remaining parameters, while the scenario of [774] used N = 58 and
M = 60, with B specified to 11 decimal places.
70
We thank Michele Cicoli, Joe Conlon, and Fernando Quevedo for helpful discussions of
the material in this section.
5.5 Inflating with K
ahler Moduli 261
but these corrections are difficult to compute. The only explicit results
available are those obtained by Berg, Haack, and Kors for N = 1 compact-
ifications on the toroidal orientifold T 6 /(Z2 Z2 ) [322, 323], cf. eqs. (3.102)
and (3.103). Some progress has been made on extending these results to
general Calabi-Yau manifolds [324, 782], and a specific functional form of
the string loop corrections has been conjectured by Berg, Haack, and Pajer
in [324] (see also [325] for a general discussion). Recall from 3.3.3 that
these corrections can be separated into two types: those associated with the
exchange of closed strings with Kaluza-Klein momentum, and those associ-
ated with the exchange of strings that wind non-contractible cycles.71 The
KK
correction K(g s)
is conjectured to be
1,1 1,1
h h
KK
X 2
CiKK (, )M KK
X CiKK (, )(a
ij tj )
K(g s)
gs gs , (5.202)
V V
i=1 i=1
where aij t is some linear combination of the two-cycle size moduli tj . The
j
W
conjectured result for K(g s)
is
W
2
X CiW (, )M X CiW (, )
W
K(g s)
, (5.203)
i
V
i (bij tj )V
where the two-cycle bij tj corresponds to the curve of intersection of two D7-
branes (see [324] for details). The unknown complex structure dependence
has been absorbed into the functions CiKK (, ) and CiW (, ).
Since we
71
The cycles in question are one-cycles within the curve of intersection of two D7-branes:
the parent Calabi-Yau manifold does not have any non-contractible (non-torsion) one-
cycles.
262 5 Examples of String Inflation
assume that the complex structure moduli are stabilized at higher energies,
we will treat them as constants, focusing on the dependence on the Kahler
moduli.
The supergravity approximation holds when ti 1, in which case
KK
X ti 1
K(g s)
> K(0 ) . (5.204)
V V
i
One might then worry that at large volume the gs corrections (5.202) and
(5.203) overwhelm the 0 corrections K(0 ) and drastically change the vac-
uum structure, to say nothing of the inflationary phenomenology. However,
what happens is a bit more subtle [324, 782]. Although the gs corrections
dominate over the 0 corrections in the Kahler potential, they cancel to a
certain degree in the scalar potential, so that the dominant contribution to
the scalar potential actually comes from the 0 corrections. This important
phenomenon is called extended no-scale structure [782]. It arises because
KK
K(g s)
is a homogeneous function, of degree 2, in the two-cycle volumes ti
[782].
To illustrate extended no-scale structure, we consider an example with a
single modulus [30]. Schematically, we can write the Kahler potential as
K = 2 ln(V) + . (5.205)
V V
Taking the superpotential to be a constant, W = W0 , we find
W02
1 1
V = 3 0 + + 0 + + 3/2 . (5.206)
V
The first zero in (5.206) corresponds to the famous no-scale structure, while
the second vanishing contribution (namely, 0 ) is the consequence of
what we just referred to as extended no-scale structure.We see that the
leading gs contribution to the scalar potential scales as 1/ , and is smaller
than the leading 0 contribution proportional to . The gs contribution to
the scalar potential is therefore smaller than naively expected. Even so, we
will find that gs corrections can still make dangerously large contributions
to the inflaton potential.
Extended no-scale structure can be understood from an alternative point
of view [782]. In the low-energy effective field theory, we can interpret the
gs contribution as the one-loop Coleman-Weinberg potential
" !#
4 2 2 4 M2
VCW ' 0 + STr(M ) + STr M ln , (5.207)
2
5.5 Inflating with K
ahler Moduli 263
where
2
Mpl 2 Mpl
= MKK ' , STr(M ) ' , (5.208)
V 2/3 V2
and STr denotes the supertrace.72 The first term in (5.207) vanishes by
supersymmetry. Substituting (5.208) into (5.207), we get
1 1 1
VCW ' 0 + + , (5.209)
V 8/3
V 10/3
V4
We will look at the part of the moduli space satisfying the hierarchies b
s . The field b then determines the overall volume, while and s
are blow-up cycles.
Inflaton potential.We first assume that string loop corrections can be ig-
nored, so that the K
ahler potential is given by (5.200). For the superpoten-
tial, we take
W = W0 + A ea T + As eas Ts . (5.212)
72 2 P 2s 2
The supertrace is defined by STr(M ) s (2s + 1)(1) Tr(Ms ), where s is the spin
2
and Ms is the matrix of masses squared for particles of spin s. Note that bosonic and
fermionic contributions enter with opposite sign.
264 5 Examples of String Inflation
stabilized
volume cycle
inflationary
blow-up cycle
stabilized
blow-up cycle
3/2
This structure stabilizes the moduli73 b and s , with V b eas s and
s 1. Integrating out b and s leads to the potential [620]74
e2a ea
!
2
V = W0 a b +c 3 , (5.213)
V V2 V
73
The axionic partner of the inflaton is not necessarily stabilized, and allowing
to have a nonvanishing initial velocity leads to the rich dynamics known as roulette
inflation [784].
74
This corrects a misprint in the corresponding formula (41) in [30].
75
See [785] for numerical evidence supporting the validity of the single-field approximation.
5.5 Inflating with K
ahler Moduli 265
ln(V)3/4
a h i O(1) ln(V) hi O(1) . (5.216)
V 1/2
Writing = hi + and using V 1, we find
' V0 1 1 e2 ,
V () (5.217)
where76
1 c1 V 5/3 hi4/3 O(V ln(V)) , (5.218)
4
2 c2 V 2/3 hi1/3 = O(V 1/2 ln(V)1/4 ) . (5.219)
3
Eta problem from string loops.The region of interest for inflation is given
by
V 2/3 4/3 1 and 1, (5.220)
where the first condition renders the potential (5.214) exponentially flat,
while the second ensures that b , i.e. that the inflationary blow-up
3/2
cycle makes a negligible contribution to the overall volume V b .
However, at this point one should remember that we have not yet included
string loop corrections. Using (5.205), we can estimate the string loop
correction to the inflaton potential,
1 1
V(gs ) 3 2/3
. (5.221)
V V 10/3
The associated correction to the parameter is
V(g00s ) 1 V
8/3 1/3 2 , (5.222)
V0 V
where we are still using units with Mpl 1. Using h i and inserting
(5.216) in (5.222), we find that
V
a2 1. (5.223)
ln(V)2
Thus, the leading string loop corrections to the Kahler potential even
after incorporating the cancellation of extended no-scale structure lead
to parametrically large values of .
76
Our volume scalings of 1 and 2 differ from [31].
266 5 Examples of String Inflation
77 (1,1,2,2,6)
The simplest example is a Calabi-Yau hypersurface in P4 see [786] for details.
5.5 Inflating with K
ahler Moduli 267
W = W0 + As eas Ts . (5.226)
The scalar potential is then
s 2as s 02 1 .
V = a2s A2s e as As W0 s eas s + W (5.227)
V V V3
The potential (5.227) depends only on s and V, which are stabilized at
s gs1 and V W0 s eas s . This leaves a flat direction in the (1 , 2 )
plane namely, the direction along which V remains constant. This flat
direction is plausibly lifted by string loop corrections to the Kahler poten-
tial. The main idea of fibre inflation is that these quantum corrections will
provide the leading (non-constant) terms in the inflaton potential. Before
proceeding, we must emphasize that the string loop corrections in question,
eqs. (5.202) and (5.203), are those conjectured in [324] (see also [325, 782])
as generalizations of the explicit computations of [322, 323] for the toroidal
orientifold T 6 /(Z2 Z2 ). The viability of fibre inflation rests on the spe-
cific form assumed in [324], and it would be valuable to obtain more direct
and detailed understanding of quantum corrections to the Kahler potential.
Without further apologies, the potential from the conjectured string loop
corrections is
!
W02 gs2 1 gs2 1
V(gs ) = 2 a 2 b +c 2 , (5.228)
V 1 1 V V
where a, b and c are unknown order-one constants. This fixes the fiber mod-
ulus 1 at 1 gs4/3 V 2/3 . An inflationary phase can arise if 1 is displaced
far from this minimum, i.e. if the K3 fiber is initially large compared to the
base, and then relaxes to smaller values.
As a simple first step, we suppose that s and V remain fixed at their
minima while 1 evolves, and can be integrated out. The resulting single-
field potential takes the form
4 /3 1 4/3 C 2/3
V () = V0 1 e + e + e , (5.229)
3 3 3
2 I II III IV
0 2 4 6 8 10 12
Fig. 5.14. Sketch of the potential for fibre inflation (figure adapted from [30]).
The phenomenologically viable inflationary regime is the gray shaded region II. The
slow-roll conditions are also satisfied in region III, but the spectrum of fluctuations
is blue.
78
The slow-roll conditions are also satisfied in region III, but constraints on the spectral
index are violated there; see 5.5.6.
5.5 Inflating with K
ahler Moduli 269
only on V, and by the string loop corrections (5.202) and (5.203), which
enter in (5.228) and lift all flat directions. Corrections from higher string
loops, and at higher order in 0 , have not been computed, but are suppressed
by additional (possibly fractional) powers of gs and V 1 .
To understand whether all such corrections can be neglected, we recall
that the remarkable resilience of LVS rests in large part on the fact that
at exponentially large V, any unknown or unwanted corrections that are
suppressed by any reasonable power (including a fractional power) of V are
effectively negligible. As an example, a possible higher-derivative correction
to the ten-dimensional action at order (0 )4 would on dimensional grounds
be suppressed compared to the leading term (3.100) by a factor V 1/3 . In
traditional LVS constructions, V 1/3 is a very small number, justifying the
omission of higher-derivative terms in ten dimensions.79
In fibre inflation, however, reproducing the normalization of the scalar
power spectrum compels the volume to be modest in size. This follows be-
cause the potential (5.231) has only one free parameter: the scale is set by
V0 V 10/3 and the slow-roll parameter is not parametrically adjustable.
For the benchmark parameters given in [43], the scalar power spectrum has
the right amplitude for V 1700. Neglecting terms suppressed by integer
powers of V is clearly safe, but suppressions by V 1/3 are marginal, partic-
ularly in cases where the dimensionless prefactor is entirely unknown.80
Higher-loop corrections are a potentially important issue in a model driven
by one-loop corrections. To understand higher-loop corrections in fibre in-
flation, we consider the limit 1 at constant V, corresponding to a K3
fiber that is large compared to the base P1 , which becomes singular in the
limit. The geometric singularity is reflected in a (power law) divergence in
the one-loop corrections involving 2 , which vanishes in the large fiber limit
[43]. An obvious concern is that higher-loop corrections to the Kahler po-
tential will become important in this regime. However, it was shown in [43]
that slow-roll inflation also breaks down at small base volume, at a value
of the base volume that is large enough so that higher-loop corrections are
still small. As a result, higher-loop corrections are argued to be negligible
during the inflationary phase.
Finally, the authors of [43] have argued that fibre inflation is robust be-
cause of a hidden symmetry that emerges in the limit of infinite volume.
In four-dimensional terms, the problematic corrections to the inflaton po-
tential are suppressed by powers of V, and vanish in the decompactification
limit V . This limit enjoys additional symmetries, most notably ten-
79 0
See [366] for the details of the expansion in LVS.
80 0
However, because the leading correction (3.100) does not depend on the inflaton, one
0
might conjecture that subleading corrections are likewise inflaton-independent, and
hence unimportant. We thank Michele Cicoli for discussions of this point.
270 5 Examples of String Inflation
81
Poly-instantons should not be confused with multi-instantons, which are well-known in
field theory and correspond in string theory to multiple Euclidean branes wrapping the
same cycle.
82
Precisely this structure arises in axion monodromy inflation through instanton correc-
tions to the holomorphic gauge coupling function f , cf. eq. (5.171).
5.5 Inflating with K
ahler Moduli 271
83
The racetrack is unrelated to the existence of poly-instanton effects: it is a further
model-building requirement. By introducing another adjustable parameter, the race-
track superpotential allows one to evade constraints that arise in a single-condensate
model [621, 783].
272 5 Examples of String Inflation
corrections to the K ahler potential that were crucial in 5.5.3 and 5.5.4. In
[621], it is argued that because D7-branes only wrap s , not 1 or 2 , with
only Euclidean D3-branes wrapping 1 , one does not expect open string
loop corrections that depend on the inflaton 1 . However, as we remarked
in 5.5.3, it has not actually been shown that dangerous open string loop
corrections are absent in this setting. Instead, a fair summary is that the
calculation of [322, 323] that led to the conjecture [324] is not immediately
applicable, and no first-principles computation of the quantum corrections
has been presented. The absence of (a certain sort of) quantum corrections
to an unprotected quantity such as the Kahler potential would be quite
striking, and further investigation of this point is warranted.
In addition to corrections from loops of open strings ending on D7-branes,
the Kahler potential can also be corrected by loops of closed strings. This
quantum correction was estimated in [621], where it was found that closed
string loops can significantly affect the shape of the inflaton potential. The
size of the effect depends on an undetermined amplitude Cloop that depends
on the complex structure moduli, and may be assumed to be of order unity
in generic situations. In [621], it was assumed that for appropriate choices of
flux one has Cloop . 0.1, in which case the loop corrections can be neglected.
5.5.6 Phenomenology
In the truncation to a single-field description, the models of inflation in
LVS described in this section can all be written in terms of the approximate
potential
V () V0 1 1 e2 . (5.239)
1 2
' 1 22 e2 and ' . (5.240)
2 22
This class of models therefore satisfies and hence (2.44) becomes
ns ' 1 + 2 . (5.241)
Given ns , one predicts the tensor-to-scalar ratio:
2 2 ns =0.96 3 103
r' (ns 1) . (5.242)
22 22
This prediction for r depends on the parameter 2 , which differs for the
different classes of K
ahler moduli inflation scenarios:
5.5 Inflating with K
ahler Moduli 273
r 105 . (5.244)
84
For the same reason, in 5.1 we did not analyze the phenomenology that would arise
in warped D3-brane inflation driven by a Coulomb potential with no corrections from
moduli stabilization: although these predictions are widely quoted in the literature,
they have little meaning.
274 5 Examples of String Inflation
of the simplest fibre inflation models have been constructed [566] that pro-
duce relatively large local non-Gaussianity from modulated reheating [170
loc
172, 256]: fNL O(few). Both possibilities are strongly constrained by the
Planck bound (1.72).
85
An effective field theory of dissipative inflation was constructed in [793]. Related work
on warm inflation [794] is reviewed in [795] (see also [767, 796]).
5.6 Inflating with Dissipation 275
Here, we describe the basic elements of that analysis and then apply them
to trapped inflation.86
Particle Production
We start by computing the particle production for a simple field theory
model [664, 798, 799]. The result of this computation will feed into the
dynamics of the inflationary model. Consider a scalar field coupled to the
inflaton via the interaction
1
Lint = g 2 ( 0 )2 2 . (5.246)
2
Notice that the field becomes massless at a specific point in field space,
= 0 . This is where the particles are produced. Near this point, we
can approximate the homogeneous inflaton evolution as
(t) 0 + 0 (t t0 ) , (5.247)
where k?2 g| 0 |. The evolution equation for a Fourier mode of the field
is then !
2
k 4 2
k + 3H k + + k? (t t0 ) k = 0 . (5.249)
a2
| {z }
2
k (t)
| k | > k2 . (5.250)
This occurs in the time interval |t t0 | < k?1 and for momenta k < k? .
Solving (5.249) gives the occupation number of the particles [664, 798]87
2 2
nk = ek /k?
. (5.251)
d3 k k?3
Z
n (t0 ) = n k . (5.252)
(2)3 (2)3
86
This section is based mostly on [622, 664].
87
This result assumes k? > H.
276 5 Examples of String Inflation
0.1626
0.1624
0.1622
Fig. 5.15. Decay of the inflaton velocity due to particle production (figure adapted
from [799]). The time t = 0 corresponds to the production event, the coupling is
g 2 = 0.1, and m2 V 00 .
+ 3H + V 0 = g 2 ( 0 )h 2 i , (5.254)
where88
n (t)
h 2 i . (5.255)
g| 0 |
Fig. 5.15 shows a numerical solution of eq. (5.254). We see that the inflaton
velocity decays after the production event, but then returns almost to its
initial value as the effect of the particles gets diluted away. The cosmological
evolution is affected only temporarily by the particle production.89 In other
words, a single particle production event does not lead to many e-folds
88
For a derivation of (5.255) see [664, 798] .
89
Particle production can be continuous if the inflaton is coupled to a gauge field (see
5.6.3).
5.6 Inflating with Dissipation 277
If the production events are spaced densely enough91 , then we can replace
the sum by an integral
X g 3/2 (ti ) a3 (ti ) Z t dt0 5/2 (t0 ) a3 (t0 ) 1 5/2 (t)
, (5.259)
i (2)3 a3 (t) (2)3 a3 (t) 3H (2)3
1 g 5/2 5/2
+ 3H + V 0 + =0. (5.260)
24 3 H
90
Trapped inflation was first proposed in [664, 800]. The inflationary mechanism, the
spectrum and bispectrum, and possible microphysical embeddings were systematically
analyzed in [622]. See also the related work [799].
91 2
The necessary condition is { /H , / } [622].
278 5 Examples of String Inflation
Notice the extra friction term proportional to 5/2 provided by the finite
3H||),
density of particles. Assuming slow-roll (|| and taking the
damping to be dominated by particle production (3H|| V 0 ), we find
2/5
= g 24 3 HV 0 . (5.261)
1 1 7/2 (t)
(t) . (5.266)
(3H)2 g (2)3
where is the potential slow-roll parameter (2.39) and we have dropped some
unimportant numerical factors. We see that inflation can occur ( < 1) even
for a steep potential ( > 1). The parametric scaling of the answer in (5.267)
is as expected: particle production is more efficient for larger coupling g and
smaller spacing ; both of these effects correspond to smaller for fixed
. Consistency conditions and further constraints on g and were studied
in [622].
5.6 Inflating with Dissipation 279
tx : (x, u1 , u2 ) 7 (x + 1, u1 , u2 ) (5.268)
tu1 : (x, u1 , u2 ) 7 (x M u2 , u1 + 1, u2 ) (5.269)
tu2 : (x, u1 , u2 ) 7 (x, u1 , u2 + 1) , (5.270)
where Lu1 , Lu2 , and Lx are dimensionless constants. This geometry corre-
sponds to a T 2 fibration over a circle parameterized by u1 , which we denote
by Su11 : for each value of u1 there is a T 2 in u2 and x,
ds2T 2 (u1 )
0 = L2u2 du22 + L2x (dx + M u1 du2 )2 . (5.272)
The identification (5.269) shows that the fiber T 2 at u1 = 1 is twisted by
an SL(2, Z) transformation before being glued to the fiber at u1 = 0. More
precisely, the complex structure of the torus shift by M units, i.e. 7 +M
as u1 7 u1 + 1. These equivalent tori are identified by the projection tu1 .
At M special locations around Su11 , M u1 = j Z, the tori are rectangular:
ds2T 2 ,
= L2x dy12 + L2u2 dy22 . (5.273)
0
We have defined coordinates y1 x + ju2 and y2 u2 obtained from an
SL(2, Z) transformation of x and u2 .
The configuration of interest is type IIA string theory compactified on
an orientifold of the product space N3 N 3 , with N 3 a second nilmanifold.
280 5 Examples of String Inflation
D4
D4
0 1 2
4
Z
1 2 10/3 2/3
SD4 = d x g , (5.275)
2
where
22 M Lu
2 = (2)3 gs 3 1 u31 , (5.276)
Mpl 9 L Lu2
!1/10
Ms 9 M 2
3
Lx Lu2
= . (5.277)
Mpl Mpl 4 (2)8 gs2 L Lu1
Here, we have defined L3 Lu1 Lu2 Lx . The field range can be super-
Planckian if Lu1 & Lu2 and
L3
u31 . (5.278)
M
This corresponds to moving around the S 1 many times (see fig. 5.16).
5.6 Inflating with Dissipation 281
wrapped
D4-brane
spectator branes
92
The first proposal to use self-collisions of a bubble in compact extra dimensions to
drive inflation appears in [801]. Cascades following nucleation events were discussed
in [802, 803]. A closely related scenario in which D-brane motion around the compact
cycle discharges a flux is [804].
93
This section is based on [623, 624].
94
As a simple analogy [624], one can picture the flux as a rubber sheet that wraps re-
284 5 Examples of String Inflation
Fig. 5.18. A cascade of five-form flux on dS4 S 1 . The compact S 1 and its
covering space are both shown.
97
A bubble of size z & ` would correspond to an ordinary bubble of reduced flux in dS4 ,
and would expand in dS4 without initiating a cascade.
286 5 Examples of String Inflation
98
This deceleration leads to bremsstrahlung, which is dramatically enhanced at large
[719]. The values of allowed by limits on non-Gaussianity are not large enough for the
results of [719], where an ultrarelativistic limit was assumed, to be directly applicable,
but the losses to closed string radiation during unwinding inflation may nevertheless be
significant, and deserve further study.
288 5 Examples of String Inflation
99
One way to achieve large is to fine-tune two axion decay constants to be nearly
coincident [767].
5.6 Inflating with Dissipation 289
100
This section is based mostly on [625, 797]. We thank Peter Adshead and Emil Martinec
for helpful discussions.
101
Any SU (N ) group has an SU (2) subgroup, and here we have identified the global part
of this SU (2) with the group SO(3) of spatial rotations.
290 5 Examples of String Inflation
where f C , /A , and g 1/ A . The same action arises in
phenomenological models of chromo-natural inflation [797].
0.2 2 10
Fig. 5.19. Evolution of a charged particle in two dimensions with quadratic ex-
ternal potential and a coupling to a homogeneous magnetic field. For large enough
coupling to the magnetic field, the particle experiences slow magnetic drift. (The
numerics for these figures was kindly provided by Peter Adshead.)
g
+ 3H + V, = 3 2 ( + H) . (5.294)
f
In addition to the force from the bare axion potential, the field experiences
the analogue of a magnetic drift force proportional to the coupling . For
large , the two forces balance each other and hence allow a slow evolution
of the inflaton. This is closely related to the magnetic drift phenomenon of
a charged particle coupled to a magnetic field; see fig. 5.19.
For sufficiently large , the effective action (5.293) leads to inflation. The
maximum number of e-folds that can be achieved while the axion rolls to
the minimum of its potential is found to be [797]
3
(Ne )max . (5.295)
5
Thus, successful chromo-natural inflation requires a large Chern-Simons
coupling, & O(100).
The crucial question is whether such a large coupling can be achieved
in a controlled string compactification. In fact, it is easy to see that this
cannot be achieved for a stack of D3-branes at weak coupling, as in that
case the Chern-Simons coupling is fixed by the string coupling, gs 1.
A possible alternative is a stack
R of D7-branes wrapping a four-cycle 4 ,
with Euler number (4 ) = (R), and with worldvolume gauge field
4
5.6 Inflating with Dissipation 291
`4
(4 )
= K+ gs s . (5.297)
24 V4
It appears difficult, but not impossible, to achieve & O(100) in this setting
while retaining control of the gs and 0 expansions. Further attempts to
obtain large magnetic couplings in string theory are discussed in [625].
5.6.4 Phenomenology
The study of the primordial perturbations arising in dissipative models is
comparatively new, and because realizations in string theory are also a
work in progress, a definitive characterization of the phenomenology is not
available at present. In this section, we will briefly describe some of the
more robust signatures.
Trapped inflation.The perturbations in trapped inflation, while under-
stood in some detail [622, 793], are not easily described analytically. We
therefore present only the main results, referring the reader to the original
literature for derivations [622, 793].
and the
P 2 in the number of produced particles, n
variance
2
g i i hi i ( i ). Solving (5.298) leads to the power spec-
trum of curvature perturbations [622]
2/3
H M
2R g 8/3 . (5.300)
M
H 1
ns 1 = . (5.302)
H2 4 H
For the specific example studied in [622], the tilt was found to be
ns = 0.99, but in general the tilt depends on the details of the model,
such as the shape of the potential, the density of particle production
events, and the properties of the particles that are produced.
If the power spectrum of tensors is dominated by vacuum fluctuations,
cf. eq. (1.34), then the tensor-to-scalar ratio is
2/3
8/2 HM
r=g 2 . (5.303)
Mpl M
equil M2
fNL ' . (5.304)
H2
The bispectrum for trapped inflation still satisfies the single-field con-
sistency condition [116, 117], as proved in [819].
102
In the case of p = 6 the Lorentz factor exceeds the limit of eq. (5.135), while p = 7 and
p = 8 are clearly incompatible with metastable compactification [623].
6
Conclusions and Outlook
Our mistake is not that we take our theories too seriously, but that we do
not take them seriously enough. It is always hard to realize that these num-
bers and equations we play with at our desks have something to do with the
real world. Even worse, there often seems to be a general agreement that
certain phenomena are just not fit subjects for respectable theoretical and
observational effort.
Steven Weinberg, on the Big Bang model [821].
294
6 Conclusions and Outlook 295
completion of gravity.
A striking feature of present observations is the extraordinary simplicity
of the primordial curvature fluctuations, which are approximately Gaussian,
adiabatic, and nearly scale-invariant. In contrast, the ultraviolet comple-
tions presented in this book are complex, involving many interacting fields
and a landscape of quantized parameters. Should the simple observations
be read as evidence against complicated models of inflation in string the-
ory? We do not believe so: although the simplicity of the data motivates
considering simple effective theories of inflation, it does not constrain the
ultraviolet completions in the same way. As an analogy, the Fermi theory
of beta decay is far simpler in terms of a counting of parameters than
the Standard Model, but is merely a low-energy effective description. In-
deed, the whole point in using effective theories is that they are simpler to
use than their ultraviolet completions. Even so, it remains important to
understand whether the simplicity of the data can emerge from the appar-
ent complexity of the ultraviolet completion: one should determine which
details of the short-distance physics decouple and which leave subtle traces
in the data.
We have largely avoided discussing deep issues involving the initial con-
ditions for inflation, including the global view of eternal inflation [822825],
the associated measure problem [826833], and the geodesic incompleteness
of inflation in the past [834]. String theory has inspired several compelling
approaches to these questions, but no complete solutions have been ad-
vanced. Many authors have noted that these unresolved problems threaten
the predictivity of the inflationary paradigm. Of course, once an inflation-
ary phase begins in a particular region of field space, clear and specific
predictions do emerge. On the other hand, it is an important open problem
to determine the relative probabilities of different inflationary models in a
broader setting. More generally, deriving specific predictions from the string
landscape as a whole, rather than from individual models, is a distant goal
that could require a new approach to the measure problem.
In closing, we would like to emphasize that the study of inflation in string
theory has advanced to a stage where a properly-constructed model can be
falsified: indeed, many models have already been falsified by recent observa-
tions, while others are under observational pressure. Proper construction of
models of string inflation, however, is a subtle art. We have railed against
incorrect predictions rooted in oversimplified effective theories, and have
catalogued the pitfalls in attempts to compute observational signatures.
Our hope is that the reader will use the ideas and techniques presented here
to derive predictions that illuminate the history of the universe and shed
light on the nature of quantum gravity.
References
297
298 References
[92] C. Blake et al., The WiggleZ Dark Energy Survey: Mapping the
Distance-Redshift Relation with Baryon Acoustic Oscillations,
Mon.Not.Roy.Astron.Soc. 418 (2011) 17071724, arXiv:1108.2635
[astro-ph.CO].
[93] F. Beutler et al., The 6dF Galaxy Survey: Baryon Acoustic Oscillations
and the Local Hubble Constant, Mon.Not.Roy.Astron.Soc. 416 (2011)
30173032, arXiv:1106.3366 [astro-ph.CO].
[94] D. H. Weinberg et al., Observational Probes of Cosmic Acceleration,
arXiv:1201.2434 [astro-ph.CO].
[95] J. Bond, G. Efstathiou, and M. Tegmark, Forecasting Cosmic Parameter
Errors from Microwave Background Anisotropy Experiments,
Mon.Not.Roy.Astron.Soc. 291 (1997) L33L41, arXiv:astro-ph/9702100
[astro-ph].
[96] M. Zaldarriaga, D. N. Spergel, and U. Seljak, Microwave Background
Constraints on Cosmological Parameters, Astrophys.J. 488 (1997) 113,
arXiv:astro-ph/9702157 [astro-ph].
[97] S. Das et al., The Atacama Cosmology Telescope: A Measurement of the
Cosmic Microwave Background Power Spectrum at 148 and 218 GHz from
the 2008 Southern Survey, Astrophys.J. 729 (2011) 62, arXiv:1009.0847
[astro-ph.CO].
[98] J. Sievers et al., The Atacama Cosmology Telescope: Cosmological
Parameters from Three Seasons of Data, arXiv:1301.0824
[astro-ph.CO].
[99] R. Keisler et al., A Measurement of the Damping Tail of the Cosmic
Microwave Background Power Spectrum with the South Pole Telescope,
Astrophys.J. 743 (2011) 28, arXiv:1105.3182 [astro-ph.CO].
[100] K. Story et al., A Measurement of the Cosmic Microwave Background
Damping Tail from the 2500-Square-Degree SPT-SZ Survey,
arXiv:1210.7231 [astro-ph.CO].
[101] Planck Collaboration, P. Ade et al., Planck 2013 Results. XVII.
Gravitational Lensing by Large-Scale Structure, arXiv:1303.5077
[astro-ph.CO].
[102] Planck Collaboration, P. Ade et al., Planck 2013 Results. XXIII. Isotropy
and Statistics of the CMB, arXiv:1303.5083 [astro-ph.CO].
[103] D. N. Spergel and M. Zaldarriaga, CMB Polarization as a Direct Test of
Inflation, Phys.Rev.Lett. 79 (1997) 21802183, arXiv:astro-ph/9705182
[astro-ph].
[104] S. Dodelson, Coherent Phase Argument for Inflation, AIP Conf.Proc.
689 (2003) 184196, arXiv:hep-ph/0309057 [hep-ph].
[105] A. Albrecht, D. Coulson, P. Ferreira, and J. Magueijo, Causality and the
Microwave Background, Phys.Rev.Lett. 76 (1996) 14131416,
arXiv:astro-ph/9505030 [astro-ph].
304 References
[495] G. Shiu and J. Xu, Effective Field Theory and Decoupling in Multi-Field
Inflation: An Illustrative Case Study, Phys.Rev. D84 (2011) 103509,
arXiv:1108.0981 [hep-th].
[496] A. Achucarro, J.-O. Gong, S. Hardeman, G. A. Palma, and S. P. Patil,
Features of Heavy Physics in the CMB Power Spectrum, JCAP 1101
(2011) 030, arXiv:1010.3693 [hep-ph].
[497] S. Cremonini, Z. Lalak, and K. Turzynski, Strongly Coupled Perturbations
in Two-Field Inflationary Models, JCAP 1103 (2011) 016,
arXiv:1010.3021 [hep-th].
[498] G. Villadoro and F. Zwirner, N = 1 Effective Potential from Dual
Type-IIA D6/O6 Orientifolds with General Fluxes, JHEP 0506 (2005)
047, arXiv:hep-th/0503169 [hep-th].
[499] O. DeWolfe, A. Giryavets, S. Kachru, and W. Taylor, Type IIA Moduli
Stabilization, JHEP 0507 (2005) 066, arXiv:hep-th/0505160 [hep-th].
[500] M. Ihl and T. Wrase, Towards a Realistic Type IIA T 6 /Z4 Orientifold
Model with Background Fluxes. Part 1. Moduli Stabilization, JHEP 0607
(2006) 027, arXiv:hep-th/0604087 [hep-th].
[501] M. P. Hertzberg, S. Kachru, W. Taylor, and M. Tegmark, Inflationary
Constraints on Type IIA String Theory, JHEP 0712 (2007) 095,
arXiv:0711.2512 [hep-th].
[502] R. Flauger, S. Paban, D. Robbins, and T. Wrase, Searching for Slow-Roll
Moduli Inflation in Massive Type IIA Supergravity with Metric Fluxes,
Phys.Rev. D79 (2009) 086011, arXiv:0812.3886 [hep-th].
[503] M. P. Hertzberg, M. Tegmark, S. Kachru, J. Shelton, and O. Ozcan,
Searching for Inflation in Simple String Theory Models: An Astrophysical
Perspective, Phys.Rev. D76 (2007) 103521, arXiv:0709.0002
[astro-ph].
[504] J. M. Maldacena, The Large N Limit of Superconformal Field Theories
and Supergravity, Adv.Theor.Math.Phys. 2 (1998) 231252,
arXiv:hep-th/9711200 [hep-th].
[505] A. Achucarro, J.-O. Gong, S. Hardeman, G. A. Palma, and S. P. Patil,
Mass Hierarchies and Non-Decoupling in Multi-Scalar Field Dynamics,
Phys.Rev. D84 (2011) 043502, arXiv:1005.3848 [hep-th].
[506] A. Achucarro et al., Heavy Fields, Reduced Speeds of Sound and
Decoupling During Inflation, Phys.Rev. D86 (2012) 121301,
arXiv:1205.0710 [hep-th].
[507] S. Hardeman, J. M. Oberreuter, G. A. Palma, K. Schalm, and T. van der
Aalst, The Everpresent Eta Problem: Knowledge of All Hidden Sectors
Required, JHEP 1104 (2011) 009, arXiv:1012.5966 [hep-ph].
[508] L. Randall and R. Sundrum, Out of this World Supersymmetry Breaking,
Nucl.Phys. B557 (1999) 79118, arXiv:hep-th/9810155 [hep-th].
References 329
[588] T. Damour and A. Vilenkin, Gravitational Wave Bursts from Cusps and
Kinks on Cosmic Strings, Phys.Rev. D64 (2001) 064008,
arXiv:gr-qc/0104026 [gr-qc].
[589] T. Damour and A. Vilenkin, Gravitational Radiation from Cosmic
(Super)strings: Bursts, Stochastic Background, and Observational
Windows, Phys.Rev. D71 (2005) 063510, arXiv:hep-th/0410222
[hep-th].
[590] H. Motohashi and T. Suyama, Detecting Cosmic String Passage through
the Earth by Consequent Global Earthquake, arXiv:1305.6676
[astro-ph.CO].
[591] X. Martin and A. Vilenkin, Gravitational Radiation from Monopoles
Connected by Strings, Phys.Rev. D55 (1997) 60546060,
arXiv:gr-qc/9612008 [gr-qc].
[592] L. Leblond, B. Shlaer, and X. Siemens, Gravitational Waves from Broken
Cosmic Strings: The Bursts and the Beads, Phys.Rev. D79 (2009) 123519,
arXiv:0903.4686 [astro-ph.CO].
[593] H. B. Nielsen and P. Olesen, Vortex Line Models for Dual Strings,
Nucl.Phys. B61 (1973) 4561.
[594] E. Witten, Cosmic Superstrings, Phys.Lett. B153 (1985) 243.
[595] N. T. Jones, H. Stoica, and S. H. Tye, Brane Interaction as the Origin of
Inflation, JHEP 0207 (2002) 051, arXiv:hep-th/0203163 [hep-th].
[596] S. Sarangi and S. H. Tye, Cosmic String Production towards the End of
Brane Inflation, Phys.Lett. B536 (2002) 185192, arXiv:hep-th/0204074
[hep-th].
[597] N. T. Jones, H. Stoica, and S. H. Tye, The Production, Spectrum and
Evolution of Cosmic Strings in Brane Inflation, Phys.Lett. B563 (2003)
614, arXiv:hep-th/0303269 [hep-th].
[598] K. Becker, M. Becker, and A. Krause, Heterotic Cosmic Strings,
Phys.Rev. D74 (2006) 045023, arXiv:hep-th/0510066 [hep-th].
[599] A. Sen, SO(32) Spinors of Type I and Other Solitons on Brane/Anti-brane
Pair, JHEP 9809 (1998) 023, arXiv:hep-th/9808141 [hep-th].
[600] G. Dvali and A. Vilenkin, Formation and Evolution of Cosmic D-Strings,
JCAP 0403 (2004) 010, arXiv:hep-th/0312007 [hep-th].
[601] E. J. Copeland, R. C. Myers, and J. Polchinski, Cosmic F and D-Strings,
JHEP 0406 (2004) 013, arXiv:hep-th/0312067 [hep-th].
[602] M. G. Jackson, N. T. Jones, and J. Polchinski, Collisions of Cosmic F and
D-Strings, JHEP 0510 (2005) 013, arXiv:hep-th/0405229 [hep-th].
[603] J. A. Harvey and A. Strominger, The Heterotic String is a Soliton,
Nucl.Phys. B449 (1995) 535552, arXiv:hep-th/9504047 [hep-th].
[604] J. H. Schwarz, An SL(2, Z) Multiplet of Type IIB Superstrings,
Phys.Lett. B360 (1995) 1318, arXiv:hep-th/9508143 [hep-th].
References 335
[686] H. Firouzjahi, L. Leblond, and S.-H. Henry Tye, The (p, q) String Tension
in a Warped Deformed Conifold, JHEP 0605 (2006) 047,
arXiv:hep-th/0603161 [hep-th].
[687] C. Burgess, J. M. Cline, and M. Postma, Axionic D3-D7 Inflation, JHEP
0903 (2009) 058, arXiv:0811.1503 [hep-th].
[688] P. S. Aspinwall and R. Kallosh, Fixing all Moduli for M-theory on
K3 K3, JHEP 0510 (2005) 001, arXiv:hep-th/0506014 [hep-th].
[689] P. Binetruy, G. Dvali, R. Kallosh, and A. Van Proeyen, Fayet-Iliopoulos
Terms in Supergravity and Cosmology, Class.Quant.Grav. 21 (2004)
31373170, arXiv:hep-th/0402046 [hep-th].
[690] Z. Komargodski and N. Seiberg, Comments on the Fayet-Iliopoulos Term
in Field Theory and Supergravity, JHEP 0906 (2009) 007,
arXiv:0904.1159 [hep-th].
[691] K. R. Dienes and B. Thomas, On the Inconsistency of Fayet-Iliopoulos
Terms in Supergravity Theories, Phys.Rev. D81 (2010) 065023,
arXiv:0911.0677 [hep-th].
[692] R. Gwyn, M. Sakellariadou, and S. Sypsas, Theoretical Constraints on
Brane Inflation and Cosmic Superstring Radiation, JHEP 1109 (2011)
075, arXiv:1105.1784 [hep-th].
[693] Z. Komargodski and N. Seiberg, Comments on Supercurrent Multiplets,
Supersymmetric Field Theories and Supergravity, JHEP 1007 (2010) 017,
arXiv:1002.2228 [hep-th].
[694] M. Berg, M. Haack, and B. Kors, On the Moduli Dependence of
Nonperturbative Superpotentials in Brane Inflation,
arXiv:hep-th/0409282 [hep-th].
[695] J. Garcia-Bellido, R. Rabadan, and F. Zamora, Inflationary Scenarios from
Branes at Angles, JHEP 0201 (2002) 036, arXiv:hep-th/0112147
[hep-th].
[696] R. Blumenhagen, B. Kors, D. Lust, and T. Ott, Hybrid Inflation in
Intersecting Brane Worlds, Nucl.Phys. B641 (2002) 235255,
arXiv:hep-th/0202124 [hep-th].
[697] M. Gomez-Reino and I. Zavala, Recombination of Intersecting D-branes
and Cosmological Inflation, JHEP 0209 (2002) 020,
arXiv:hep-th/0207278 [hep-th].
[698] A. Avgoustidis, D. Cremades, and F. Quevedo, Wilson Line Inflation,
Gen.Rel.Grav. 39 (2007) 12031234, arXiv:hep-th/0606031 [hep-th].
[699] A. Avgoustidis and I. Zavala, Warped Wilson Line DBI Inflation, JCAP
0901 (2009) 045, arXiv:0810.5001 [hep-th].
[700] T. W. Grimm, M. Kerstan, E. Palti, and T. Weigand, On Fluxed
Instantons and Moduli Stabilisation in IIB Orientifolds and F-theory,
Phys.Rev. D84 (2011) 066001, arXiv:1105.3193 [hep-th].
References 341
[780] T. He, S. Kachru, and A. Westphal, Gravity Waves and the LHC: Towards
High-Scale Inflation with Low-Energy SUSY, JHEP 1006 (2010) 065,
arXiv:1003.4265 [hep-th].
[781] A. Linde, Y. Mambrini, and K. A. Olive, Supersymmetry Breaking due to
Moduli Stabilization in String Theory, Phys.Rev. D85 (2012) 066005,
arXiv:1111.1465 [hep-th].
[782] M. Cicoli, J. P. Conlon, and F. Quevedo, Systematics of String Loop
Corrections in Type IIB Calabi-Yau Flux Compactifications, JHEP 0801
(2008) 052, arXiv:0708.1873 [hep-th].
[783] R. Blumenhagen, X. Gao, T. Rahn, and P. Shukla, Moduli Stabilization
and Inflationary Cosmology with Poly-Instantons in Type IIB Orientifolds,
JHEP 1211 (2012) 101, arXiv:1208.1160 [hep-th].
[784] J. R. Bond, L. Kofman, S. Prokushkin, and P. M. Vaudrevange, Roulette
Inflation with K
ahler Moduli and their Axions, Phys.Rev. D75 (2007)
123511, arXiv:hep-th/0612197 [hep-th].
[785] J. J. Blanco-Pillado, D. Buck, E. J. Copeland, M. Gomez-Reino, and N. J.
Nunes, K ahler Moduli Inflation Revisited, JHEP 1001 (2010) 081,
arXiv:0906.3711 [hep-th].
[786] P. Candelas, X. De La Ossa, A. Font, S. H. Katz, and D. R. Morrison,
Mirror Symmetry for Two-Parameter Models. 1., Nucl.Phys. B416
(1994) 481538, arXiv:hep-th/9308083 [hep-th].
[787] R. Blumenhagen and M. Schmidt-Sommerfeld, Power Towers of String
Instantons for N = 1 Vacua, JHEP 0807 (2008) 027, arXiv:0803.1562
[hep-th].
[788] I. Garcia-Etxebarria and A. M. Uranga, Nonperturbative Superpotentials
Across Lines of Marginal Stability, JHEP 0801 (2008) 033,
arXiv:0711.1430 [hep-th].
[789] R. Blumenhagen, X. Gao, T. Rahn, and P. Shukla, A Note on
Poly-Instanton Effects in Type IIB Orientifolds on Calabi-Yau Threefolds,
JHEP 1206 (2012) 162, arXiv:1205.2485 [hep-th].
[790] D. L
ust and X. Zhang, Four K ahler Moduli Stabilisation in Type IIB
Orientifolds with K3-fibred Calabi-Yau Threefold Compactification, JHEP
1305 (2013) 051, arXiv:1301.7280 [hep-th].
[791] M. Cicoli, S. Downes, and B. Dutta, Power Suppression at Large Scales in
String Inflation, arXiv:1309.3412 [hep-th].
[792] F. G. Pedro and A. Westphal, Low-l CMB Power Loss in String Inflation,
arXiv:1309.3413 [hep-th].
[793] D. Lopez Nacir, R. A. Porto, L. Senatore, and M. Zaldarriaga, Dissipative
Effects in the Effective Field Theory of Inflation, JHEP 1201 (2012) 075,
arXiv:1109.4192 [hep-th].
[794] A. Berera, Warm Inflation, Phys.Rev.Lett. 75 (1995) 32183221,
arXiv:astro-ph/9509049 [astro-ph].
References 347