26
arXiv:1804.07541v1 [astro-ph.CO] 20 Apr 2018 RESCEU-05/18 Hybrid Quasi-Single Field Inflation Yi Wang a,b , Yi-Peng Wu c , Jun’ichi Yokoyama c,d,e , and Siyi Zhou a,b a Department of Physics, The Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong, P.R.China b Jockey Club Institute for Advanced Study, The Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong, P.R.China c Research Center for the Early Universe (RESCEU), Graduate School of Science, The University of Tokyo, Tokyo 113-0033, Japan d Department of Physics, Graduate School of Science, The University of Tokyo, Tokyo 113-0033, Japan and e Kavli Institute for the Physics and Mathematics of the universe (Kavli IPMU), WPI, UTIAS, The University of Tokyo, 5-1-5 Kashiwanoha, Kashiwa 277-8583, Japan (Dated: April 23, 2018) Abstract The decay of massive particles during inflation generates characteristic signals in the squeezed limit of the primordial bispectrum. These signals are in particular distinctive in the regime of the quasi-single field inflation, where particles are oscillating with masses comparable to the Hubble scale. We apply the investigation to a class of scalar particles that experience a so-called waterfall phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on the hybrid inflation scenario. With a time-varying mass, a novel shape of oscillatory bispectrum is presented as the signature of a waterfall phase transition during inflation. * Electronic address: [email protected] 1

Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

arX

iv:1

804.

0754

1v1

[as

tro-

ph.C

O]

20

Apr

201

8RESCEU-05/18

Hybrid Quasi-Single Field Inflation

Yi Wanga,b, Yi-Peng Wuc,∗ Jun’ichi Yokoyamac,d,e, and Siyi Zhoua,b

aDepartment of Physics, The Hong Kong University of Science and Technology,

Clear Water Bay, Kowloon, Hong Kong, P.R.China

bJockey Club Institute for Advanced Study,

The Hong Kong University of Science and Technology,

Clear Water Bay, Kowloon, Hong Kong, P.R.China

cResearch Center for the Early Universe (RESCEU),

Graduate School of Science, The University of Tokyo, Tokyo 113-0033, Japan

dDepartment of Physics, Graduate School of Science,

The University of Tokyo, Tokyo 113-0033, Japan and

eKavli Institute for the Physics and Mathematics

of the universe (Kavli IPMU), WPI, UTIAS,

The University of Tokyo, 5-1-5 Kashiwanoha, Kashiwa 277-8583, Japan

(Dated: April 23, 2018)

Abstract

The decay of massive particles during inflation generates characteristic signals in the squeezed

limit of the primordial bispectrum. These signals are in particular distinctive in the regime of the

quasi-single field inflation, where particles are oscillating with masses comparable to the Hubble

scale. We apply the investigation to a class of scalar particles that experience a so-called waterfall

phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based

on the hybrid inflation scenario. With a time-varying mass, a novel shape of oscillatory bispectrum

is presented as the signature of a waterfall phase transition during inflation.

∗ Electronic address: [email protected]

1

Page 2: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

I. INTRODUCTION

Inflation [1, 2] has been a leading paradigm for the primordial universe and most likely an

observable phenomenon triggered at the highest energy scale among all current experiments.

Observations for the inflationary parameters are in well agreement with the simplest scenario

where only a single scalar degree of freedom is involved. The primordial scalar mode, ζ ,

seems to be responsible to the soft breaking of the global time-translation symmetry during

inflation [3, 4]. One therefore expects that the correlation functions of ζ follow the single-

field consistency relations [5, 6], and signals from the primordial bispectrum vanishes in the

squeezed limit unless there exist new particles that decay into ζ . This is a window opened

known as the cosmological collider physics [7], see also [8–14, 16, 17].

Massive particles with a mass m around the order of the Hubble parameter H during

inflation appear to be the main target for the cosmological collider research. This is because

massive fields with m ≫ H are usually recast into an effective background, while particles

with m ≪ H often lead to generalized version of inflationary dynamics beyond the single-

field configuration. One of the characteristic signature of massive fields with m ∼ H comes

from the three-point function of ζ in the squeezed limit. In particular, an oscillatory behavior

of the non-Gaussian correlation functions can be generated by the decay of massive fields

into ζ with m > 3H/2, as a consequence of the quantum interference between massless and

massive mode functions. This type of oscillatory feature in the squeezed non-Gaussianities

can be easily derived from the spatial dilation symmetry of inflation [12].

Quasi-single field inflation1 [8, 9] is a pioneer study which resolves the imprints of massive

fields in the non-Gaussianities with m ≥ H . As a practical example, one can consider

an inflaton field slowly rolling with a turning trajectory, whose kinetic term is naturally

mixed with a massive mode in the isocurvature direction. In this scenario, the isocurvature

perturbations receive an effective mass from both the potential and the kinetic couplings to

the inflaton. Recent studies [18–21] have confirmed that the bispectrum involved with the

decay of the massive field are indeed oscillating with respect to the ratio klong/kshort, where

klong (kshort) is a larger (smaller) scale mode which exits the horizon earlier (later) in time.

The amplitude of such an oscillation is enhanced in the squeezed limit (klong/kshort → 0) if

the large effective mass comes from the field potential [21, 22], and it is enhanced in the

1 The effective single-field inflation reproduced from a multi-field configuration with heavy field masses

m ≫ H was first investigated in [15].

2

Page 3: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

equilateral limit (klong/kshort → 1) if the large effective mass is induced by the kinetic coupling

[18, 19]. These oscillatory signals are usually small due to the Boltzmann suppression factor

e−πm/H . There are a few possibilities to generate this type of oscillatory signals without any

exponential suppression. For example, from classical oscillations of inflationary fluctuations

from a sharp feature (the classical primordial standard clock) [23–25], particle production

enhanced by inflaton kinetic energy during monodromy inflation [26], and production of

heavy particles due to finite-temperature effects during warm inflation [27].

It is also interesting to seek for signals from another inflationary scenario involving massive

fields. For example, in the hybrid inflation scenario [28] an isocurvature field acquires an

effective mass by virtue of a direct coupling to the inflaton field. This effective mass square

decays with the slow rolling of the inflaton and becomes negative after some critical points

where the isocurvature field can start to roll down into a new minimum of the potential

(namely a waterfall phase transition) 2. If the negative mass term during the falling down

process is sufficiently large and the massive field comes to dominate the energy density,

the slow-roll inflation will be terminated soon after the waterfall phase transition [28]. On

the other hand, if this mass is not much greater than the Hubble parameter, this field will

induce a secondary slow-roll inflation along the direction orthogonal to the original inflation

[34, 35], in a way similar to the double-inflation scenario [36–38]. These features lead us the

curiosity on the implications of such a kind of waterfall field to the primordial spectra. We

are especially interested in the case where the waterfall field is always subdominant in the

total energy density, even after the phase transition. In this case the background dynamics

is still governed by the single-field inflation paradigm, yet once the waterfall field falls down

into the new minimum and acquires a large positive mass, the decay of isocurvature field

perturbations can generate oscillating features in the bispectum of ζ through the same

mechanism as that of the quasi-single field inflation [7, 9].

In this work, we consider the presence of a waterfall phase transition in a turning tra-

jectory by introducing a coupling as φ2σ2/2 between the inflaton φ and the waterfalling

isocurvature field σ from the hybrid inflation [28] to the simplest quasi-single field inflation

model with a constant turning radius [9]. The hybrid potential of our model can be schemat-

ically illustrated by Fig. 1. We show that both the hybrid inflation and the constant turn

2 Phase transitions of an isocurvature field can also be realized by its direct coupling with gravity, see

[32, 33].

3

Page 4: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

inflation scenario get benefits from each other as follows:

• A turning trajectory in the inflationary direction naturally breaks the degeneracy of the

local-minimal states in the isocurvature direction, even if these minimal states share

the same vacuum expectation value. This is true when the waterfall field is accelerated

by a non-vanishing centrifugal force due to the inflaton velocity. For example, in the

hybrid inflation [28] with a double-well potential ∝ (σ2 − v2)2 the rate of σ to fall

down into the minima at ±v now depends on the initial conditions at the time of

phase transition. As a result, the rate of multi-stream inflation with domain wall

formation [29–31] in this scenario can be controlled by the model parameters.

• The oscillating signature in the ζ-bispectrum can be efficiently amplified in the

squeezed limit (klong/kshort → 0), when comparing with the results of the constant-turn

inflation [18, 19]. This is due to the fact that a waterfall phase transition driven by

the unstable part of the potential along the isocurvature direction in general leads to

a period with tachyonic instability where isocurvature field perturbations are growing

on superhorizon scales.

• A waterfall phase transition can act as a trigger for classical oscillations of the isocur-

vature field (see Fig. 1). The small oscillations of the isocurvature field can result

in observable features that are sensitive to the background expansion rate of the pri-

mordial universe. These oscillating features in the primordial spectra are recognized

as the classical standard clock signals, which are useful for testing inflation and its

alternative scenarios [23–25].

This paper is organized as the follows. In Section II, we introduce a model which in-

corporates the constant turn inflation with the waterfall potential from the hybrid inflation

scenario. We study the classical field evolution via the stochastic approach in Section III.

In Section IV, we outline our approach for the quantum field perturbations in this scenario,

and the numerical results for the two-point and three-point correlation functions are given

in Section V. Finally, we give summary and discussions in Section VI.

4

Page 5: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

σ

FIG. 1. Left panel: Illustration of the hybrid-inflation potential with a turning minimum. Right

panel: A numerical example for σ(N) that generates an ocsillating trajectory due to the waterfall

phase transition. The dashed line depicts the evolution of the positive branch of σmin.

II. HYBRID INFLATION WITH A TURN

We construct a theory based on the quasi-single field inflation [9] with a turning trajectory.

It is convenient to apply the polar-coordinate representation to decompose the inflaton φ by

the tangential mode θ = φ/R and the isocurvature field (or the isocurvaton) by the radial

mode σ. The action is given by

Sqs =

d4x√−g

[

−1

2(R + σ − σ∗

0)2(∂θ)2 − 1

2(∂σ)2 − V (θ, σ)

]

, (1)

where R and σ∗0 are constants with the same dimension of the mass. We can rephrase the

action by the usual representation of the inflaton field φ = Rθ, which has the same dimension

as σ, and (1) is rewritten as

Sqs =

d4x√−g

[

−1

2(1 +

σ − σ∗0

R)2(∂φ)2 − 1

2(∂σ)2 − V (φ, σ)

]

. (2)

The kinetic coupling (σ − σ∗0)

2(∂φ)2/R2 is assumed to be perturbatively small.

Both θ and σ exhibit a well-defined coherent part (or the zero-mode part) θ0 and σ0, and

5

Page 6: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

their evolution in the spatially homogeneous background are governed by [22]:

3M2pH

2 =1

2(R + σ0)

2θ20 +1

2σ20 + V (θ0, σ0), (3)

−2M2p H = (R + σ0)

2θ20 + σ20 , (4)

(R + σ0)2θ0 + 3(R + σ0)

2Hθ0 + Vθ = 0, (5)

σ0 + 3Hσ0 + Vσ = (R + σ0)θ20, (6)

where R ≡ R−σ∗0 , Vθ ≡ ∂V/∂θ (Vσ ≡ ∂V/∂σ), and (R+σ0)θ

20 features the centrifugal force

applied to the isocurvaton σ (which is nothing but the waterfall field in the current study).

The time evolution of σ0 plays a fundamental role in our scenario, but we shall keep σ0 ≪ R

and σ20 ≪ R2θ20 so that θ0 dominates the inflationary dynamics and observables.

A phase transition of σ0 from the initial expectation value σ∗0 may be realized by direct

couplings between θ and σ in the potential. As a simple example, we consider the symmetry-

breaking scenario driven by the rolling of θ as [28]

V (θ, σ) = Vsr(θ) +λ

4

[

(σ − σ∗0)

2 − v2]2

+g2

2θ2(σ − σ∗

0)2, (7)

where Vsr(θ) stands for the potential of the slow-roll inflation and λ is a dimensionless

parameter. g and v are parameters with the same dimension of the mass. In terms of

φ = Rθ, one can identify the dimensionless coupling constant g = g/R from the interaction

term g2θ2(σ−σ∗0)

2 ≡ g2φ2(σ−σ∗0)

2. The critical value of phase transition in the isocurvature

direction is θ2c = λv2/g2 when σ = σ∗0 . There are two interesting regimes in the theory:

1. If Vsr(θc) ≪ λv4/4, the waterfall field comes to dominate the energy density of the

universe after phase transition, and the slow-roll inflation due to Vsr(θ) can be termi-

nated by a rapid waterfall of σ [28] or be replaced by a secondary inflation driven by

the slow rolling of σ [35], depending on the value of the critical mass m2c ≡ g2θ2c = λv2.

We consider this regime as the usual hybrid inflation scenario.

2. If Vsr(θc) ≫ λv4/4, the slow-roll inflation due to θ continues after the phase transition,

but the inflationary trajectory is shifted by the rolling of the isocurvaton in the radial

direction. This regime can be cast into a broader class of the quasi-single field inflation

[9], provided that m2c ∼ H2. If m2

c ≫ H2, namely σ acquires a negative effective mass

much greater than the hubble parameter, the transition of σ finishes instantly and the

predictions of the theory converge to that of the single-field inflation.

6

Page 7: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

We therefore restrict the current study to the second regime.

A. Inflation with a constant turn

Let us put σ very close to σ∗0 at the beginning. When θ ≫ θc, its mass reads m2

σ ≈g2θ2 ≫ m2

c . If we consider that m2c & O(H2), then σ0 remains very close to σ∗

0 before the

phase transition. Thus we can expand Vσ around the minimum as

Vσ(σ) = Vσσ(σ∗0)(σ − σ∗

0) +1

2Vσσσ(σ

∗0)(σ − σ∗

0)2 + · · · , (8)

where Vσσ(σ∗0) = g2θ2 − λv2 ≫ m2

c at this stage. The equation of motion for σ0 at zeroth-

order reads

σ0 + 3Hσ0 = Rθ20, (9)

since Vσ(σ∗0) = 0.

To see the effect of a non-vanishing centrifugal force, it is convenient to consider a constant

inflaton velocity θ0. We will show that this condition holds in a generic class of slow-roll

inflation models. Ignoring for the moment the mild time-evolution of the Hubble parameter,

the solution of (9) takes the form of

σ0 = σ∗0 +

Rθ203H

t− c03H

e−3Ht, (10)

where c0 can be determined by a given initial condition but its contribution is suppressed

by the exponential factor. One can check that higher-order corrections to σ0 due to (8) are

also decaying with time. As a result, a non-zero centrifugal force leads to a deviation from

the minimum as σ0 − σ∗0 ∝ ∆t after some finite duration.

B. Dynamics of waterfall phase transition

To make a concrete discussion on the dynamics of phase transition, we redefine the

isocurvaton field into a dimensionless parameter as σ ≡ (σ0 − σ∗0)/R and we rewrite the

equation of motion (6) with respect to the number of e-folds N = ln a as

d2σ

dN2+ (3− ǫH)

dN+

[

g2θ20H2

H2(R2σ2 − v2)

]

σ =θ20H2

, (11)

7

Page 8: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

where ǫH = −H/H2 is the first slow-roll parameter. We also specify the slow-roll potential

by the chaotic inflation form

Vsr(θ) = µ4θp, (12)

which recovers the prototype scenario [28] when taking p = 2. In the single-field limit both

slow-roll parameters ǫH and ηH = ˙ǫH/(HǫH) beocme independent of µ. To sustain a large

enough e-folding number for a slow-roll inflation one requires R ≫ Mp, and the inflaton

solution is well-approximated by

θ0 = θc − pM2

p

R2θ2c(N −Nc), (13)

throughout our discussion, where Nc is the epoch when θ = θc. The potential (12) supports

a constant centrifugal force θ/H = −pM2p/(R

2θ2c ).

Without a loss of generality, we choose Nc = 0 and approximate the value of σ at the

critical point as σ(Nc) = σc ≈ θ20

3H2c

∆N , where H2c ≡ 3M2

p/(µ4θpc ) and ∆N is the e-folding

number before phase transition. Assuming that σ starts to roll down smoothly once θ reaches

the critical value, we can simplify the equation of motion (11) as

dN− ANσ = B, (14)

where we introduce two useful parameters as

A =2pg2

3H2c θc

M2p

R2, B =

p2M4p

3R4θ4c. (15)

The solution of (14) reads

σ = eAN2/2

[

σc +B

π

2AErf

(

A

2N

)]

. (16)

Since σc ≈ B∆N , it turns out that σ ≃ σceAN2/2 if ∆N ≫

π/(2A).

The waterfall ends when σ reaches the new minima at θ > θc. Taking the solution (13),

the condition Vσ = 0 with N > 0 indicates that

σmin = ±√

3H2cA

λR2N, (17)

which is evolving with time. We can estimate the epoch N = Nf at the end of the phase

transition by σ(Nf ) = σmin. The result from the solution (16) is given in Fig. 2.

8

Page 9: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

� �� ��

��

��

��

���/��

��

FIG. 2. The duration of the waterfall field to reach the positive branch of σmin since N = 0 with

respect to the critical mass m2c = g2θ2c = λv2.

III. CLASSICAL FIELD EVOLUTION

In this section we investigate the behavior of the long wavelength part of the waterfall

field σ. By imposing a classical initial value σ0 and a certain probability distribution of

the long wavelength modes, we study the evolution of such a distribution undergoing the

waterfall phase transition. In general, we identify the peak of the probability distribution

with the classical expectation value σ0(t) at a given time t. The decomposition σ → σ0+δσ,

with δσ as quantum fluctuations, becomes well-defined if the spread-out (or decoherent)

time-scale of the distribution is comparable to the duration of inflation.

The long wavelength part of σ can be treated as an auxiliary classical field, whose one-

point probability distribution function (PDF), ρ1[σ(x, t)], satisfies the Fokker-Planck equa-

tion [39, 40]∂ρ1[σ]

∂t=

1

3H

∂σ

(

ρ1[σ]∂Veff

∂σ

)

+H3

8π2

∂2ρ1[σ]

∂σ2, (18)

where the effective field potential Veff deduced from the classical equation of motion (6)

reads

Veff =1

2(g2θ20 − λv2 − ξθ20)σ

2 +λ

4σ4 − ξRθ20σ. (19)

Here the terms with θ20 arise due to the centrifugal force of the turning trajectory, and we

have moved the origin to σ = σ∗0 = 0 for convenience. We introduce a new parameter ξ

to control the magnitude of the centrifugal force only for the study in this section, and the

model (1) can be recovered by taking ξ = 1.

9

Page 10: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

FIG. 3. The evolution of the one-point probability distribution function ρ1 of the waterfall field

without centrifugal force (ξ = 0). The waterfall phase transition starts at N = 0. The initial value

σ0 = 0.5H is used. The potential parameters are fixed by g/H = 4, λ = 10−6, θc = 1, p = 1 ,

R/Mp = 10 and H/Mp = 10−5.

We adopt the analytical expression (13) for θ0 and solve ρ1 numerically by (18). We

impose an initial distribution ρ1(t0) ≡ ρ10 of the form

ρ10 =H

2π〈σ2〉0exp

[

−(σ − σ0)2

2√

〈σ2〉0

]

, (20)

where the initial variance 〈σ2〉0 is in principle a free parameter. Without a loss of generality,

we estimate the variance by 〈σ2〉0 = 3H4/(8π2m2σ0), assuming that σ is initially in an

equilibrium state with a mass m2σ = m2

σ0 [40].

To see the evolution of ρ1 in the standard hybrid inflation scenario [28], let us firstly

turn off the centrifugal force by setting ξ = 0. With an initial value σ0 6= 0, the numerical

result in Fig. 3 shows that the PDF can reach back to a normal distribution with a peak

at the origin by the time of phase transition. The PDF spreads out soon during the phase

transition as m2σ becomes very small. After the phase transition the PDF equally distributes

to the positive and the negative minima of the potential, which indicates the well-known

10

Page 11: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

FIG. 4. The evolution of the one-point probability distribution function ρ1 of the waterfall field

with a centrifugal force (ξ = 0.01). The waterfall phase transition starts at N = 0. The initial

value σ0 = −0.5H is used. The potential parameters are fixed by g/H = 4, λ = 10−6, θc = 1,

p = 1 , R/Mp = 10 and H/Mp = 10−5.

domain wall problem [28, 35].

On the other hand, if we turn on the centrifugal force by taking ξ > 0 the potential

minimum is shifted from the origin to σ > 0 by the time of phase transition. The numerical

result in Fig. 4 shows that the peak of the PDF runs into the positive region, even with

an initial value σ0 < 0, and the possibility to find σ in the negative minimum after phase

transition can be significantly suppressed. We can estimate the probability of σ to end up

in the negative minimum by

Pc =

∫ 0

−∞

ρ1c[σ]dσ

H, (21)

where ρ1 c is the one-point PDF at the critical time. To evade the domain wall problem,

we require Pc < 1/e3∆N with ∆N being the number of e-folds required to solve the horizon

problem substracted by that elapsed after the phase transition. We have checked that the

condition Pc < 1/e3∆N can be satisfied for σ0 = 0, ξ ∈ {10−3, 1} and ∆N ∼ O(40). This

implies that hybrid inflation with a turning trajectory is free from the domain wall problem

11

Page 12: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

for a wide range of initial conditions.

IV. APPROACHES FOR QUANTUM FLUCTUATIONS

We resolve the quantum fluctuations of the inflaton (δθ = δφ/R) and the waterfall

fields (δσ) by virtue of the equation of motion (EoM) approach [22], which is in particular

convenient for numerical computations. The reason is twofold. First of all, so far there is no

exact solution found for the mode function of a waterfall field, δσk, in the hybrid inflation

scenario with a time-varying mass. Although some analytical expressions for δσk can be

obtained via the WKB method [33, 42] or the Hankel function solution after specifying the

inflaton dynamics [41], these approximations capture only the dominant contributions on

limited physical scales during inflation and they might not be very useful to our purpose

since the quantum clock signals are sometimes generated by subdominant parts of the mode

functions [18, 20].

Secondly, the system inevitably becomes non-perturbative in the asymptotic future (N ≫Nc) where the inflaton-isocurvaton coupling g2θ2σ2 gives the most important contribution

to the field perturbations. To be more explicitly, let us write down the quadratic Lagrangian

for the field perturbations into

L2 =a3

2R2

[

(1 + σ)2δθ2 − 1

a2(1 + σ)2(∂iδθ)

2 + δ ˙σ2 − 1

a2(∂iδσ)

2 − (Vσσ − θ20)δσ2

]

, (22)

δL2 = 2a3R2(1 + σ)θ0δσδθ − a3RVθσδσδθ, (23)

where δθ and δσ ≡ δσ/R are decoupled in L2 but mixed in δL2.

In the interaction picture, L2 is used to define the EoM of “free fields” and δL2 is generally

cast into interactions, if it can be expanded perturbatively. This is one of the underlying

principles for the conventional in-in formalism [43–45]. In the parameter regime we are

interested, however, δL2 also plays a fundamental role in the EoM of field perturbations,

and thus the definition of free fields based on L2 becomes insufficient.

In the EoM approach [22], one can define a mixed vacuum state shared by many quantum

fields to deal with a strongly coupled system. We therefore define the mode functions with

12

Page 13: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

respect to two sets of vacuum states as

δθk = u+k ak + u+∗

k a†−k + u−k bk + u−∗

k b†−k, (24)

δσk = v+k ak + v+∗k a†−k + v−k bk + v−∗

k b†−k, (25)

where a and b are shared by δθ and δσ, and they shall satisfy the commutation relations

given by [22] as

[ak, a†−p] = [bk, b

†−p] = (2π)3δ3(k+ p), (26)

[ak, a−p] = [bk, b−p] = [ak, b−p] = [a†k, b†−p] = [ak, b

†−p] = [bk, a

†−p] = 0. (27)

Here δθ and δσ are governed by the EoMs derived from L2 together with δL2, which read

δθ′′k−(

2

z− 2σ′

1 + σ

)

δθ′k + δθk (28)

=

[

2Vθ

R2H2(1 + σ)− 2g2

θσ

(1 + σ)2

]

δσk

z+

2√3B

1 + σ

[

σ′

1 + σ

δσk

z+

δσ′k

z

]

,

δσ′′k−

2

zδσ′

k + δσk +

[

3λR2

H2σ2 + 2pg2

M2p

R2ln

(

−kckz

)

− 3B

]

δσk

z2

= −2(1 + σ)√3B

δθ′kz

− 2g2θ σδθkz2

, (29)

where z = kη is the conformal time co-moving with each mode k, and kc is the horizon scale

at the time of the phase transition. We define a dimensionless parameter g ≡ g/H . The wa-

terfall phase transition at N = − ln(−kcz/k) = − ln(−kcη) = 0 breaks the time-translation

symmetry z → cz of the mode functions, where c = k/kc. Therefore k-dependence appears

in the effective mass

M2σ(z) = 3λ

R2

H2σ2 + 3A ln

(

−z

c

)

− 3B, (30)

when using z as the time parameter, but the scale-dependence in M2σ disappears if we use

the usual conformal time η = z/k. Similarly, the EoM of the background field given by

σ′′ − 2

zσ′ +

[

λR2

H2σ2 + 3A ln

(

−z

c

)

]

σ

z2=

3B

z2, (31)

looks different for each k mode in its co-moving time z. As shown in Fig. 5, modes with

k < kc (or c < 1) experience the waterfall transition at −z = c after they have exited the

horizon at −z = 1. The k-dependence in σ and M2σ simply reflects the time dependence of

the background solution in the evolution of perturbations.

13

Page 14: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

◆ c = 0.0001

◆ c = 0.001

◆ c = 0.01

◆ c = 0.1

◆ c = 1

��-� ��-� ���� ��

������

������

������

������

������

������

-�

σ◆ c = 0.0001

◆ c = 0.001

◆ c = 0.01

◆ c = 0.1

◆ c = 1

��-� ��-� ���� ��

��

��

-�

�σ�

FIG. 5. The evolution of σ (left panel) and M2σ (right panel) with respect to the co-moving

conformal time z of various c = k/kc. The parameters are fixed by g ≡ g/H = 4, λ = 10−6, θc = 1,

p = 1 , R/Mp = 10 and H/Mp = 10−5.

The only requirement for solving the field perturbations by the EoM approach is that δθ

and δσ shall satisfy a well-posed initial condition [22]. Fortunately in the limit well inside

the horizon (−z → ∞) Eq. (28) and Eq. (29) simply reduce to

δθ′′k + δθk = 0, and δσ′′k + δσk = 0, (32)

which implies the solution of δθ ∝ e−iz and δσ ∝ e−iz. After imposing the initial condition

σ = 0, the EoM (28) and (29) coincide with the equations of the non-perturbative quasi-

single field inflation [18, 19] up to the order of 1/z. Thus the initial mode functions are

found as

u±k =

H

R√4k3

e−iz(−z)1±ipM2p/(θcR2), v±k = ±iu±

k . (33)

The slow-roll condition M2p ≪ R2 for the potential (12) implies that the two sets of mode

function u±k (and thus v±k ) are almost degenerate up to a phase factor.

14

Page 15: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

◆ c = 0.0001

◆ c = 0.001

◆ c = 0.01

◆ c = 0.1

◆ c = 1

��-� ��-� ���� ��

��-��

��-��

��-��

��-��

��-�

��-�

-�

�θ

◆ c = 0.0001

◆ c = 0.001

◆ c = 0.01

◆ c = 0.1

◆ c = 1

��-� ��-� ���� ��

��-��

��-��

��-��

��-��

��-��

��-�

-�

�σ

FIG. 6. The evolution of Pθ(c, z) (left panel) and Pσ(c, z) (right panel) with respect to the co-

moving conformal time z of various c = k/kc. The parameters are fixed by g ≡ g/H = 4, λ = 10−6,

θc = 1, p = 1 , R/Mp = 10 and H/Mp = 10−5.

V. IMPRINTS OF WATERFALL PHASE TRANSITION

A. Power spectrum

We now solve the correlation functions of δθ and δσ through the EoMs (28), (29) and

(31) from the initial states (33). The two-point correlation function for field perturbations

based on the mixed vacuum states (24) and (25) are defined as [22]:

〈δθk(z)δθp(z)〉 = (2π)3δ3(k+ p)[

u+k (z)u

+∗p (z) + u−

k (z)u−∗p (z)

]

, (34)

and for 〈δσ2〉 one simply replaces u±k by v±k , respectively. In the interaction picture, the

definition (34) includes the resummed tree-level corrections from δL2 to the free δθ field

defined by L2. The gauge invariant curvature perturbation is related to (34) by [9]

〈ζkζp〉 =H2

θ2〈δθkδθp〉 = (2π)3δ3(k+ p)

H2

θ22π2

k3Pθ(k). (35)

The results of Fig. 6 indicate that δθ is governed by two kinds of solutions. One can check

that δθ decays with a power of −z inside the horizon (−z > 1), since we have chosen a small

kinetic coupling between θ and σ, and it goes to a constant on superhorizon scales (−z < 1).

On the other hand, the evolution of δσ essentially has three phases. While δσ remains well

inside the horizon (−z ≫ 1), it is governed by the initial state solution (33) scaling as −z.

15

Page 16: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

◆ g˜= 0

◆ g˜= 1

◆ g˜= 2

◆ g˜= 3

◆ g˜= 4

����� ����� �� ����

�����

�����

�����

�����

�����

�θ

◆ IR EFT

◆ IR EFT

◆ g˜= 3

◆ g˜= 4

����� ����� �� ����

��� × ��-��

���× ��-��

� ��× ��-��

���× ��-��

�σ

FIG. 7. The value of Pθ (left panel) and Pσ (right panel) in terms of c = k/kc. The value of Pθ

is normalized with respect to the constant turn scenario without the waterfall transition (g = 0).

The dashed lines in the right panel are given by (36). The parameters are fixed by λ = 10−6,

θc = 1, p = 1 , R/Mp = 10 and H/Mp = 10−5.

Once the mass M2σ becomes important as M2

σ ≃ −z, δσ starts to oscillate. Note that δσ

grows with time at the beginning of the oscillation phase since Mσ becomes temporarily

negative. As −z → 0, σ has settled down in the new minimum (the true vacuum) given by

(17), and δσ obtain a very large mass from the evolution of σ0. In this case one may omit

all derivatives of δσ in (29) to find the mildly decaying long wavelength solution as

δσIR ≈ −2g2

H2θcσmin

M2σ

δθIR. (36)

Here δθIR ≈ H/(√2R) stands for the asymptotic value of δθk in the limit −z → 0, which

is nearly scale-independent. This kind of effective field theory approach can be applied to

both background level and perturbation level as is widely used in the previous literatures

[46–56]. By using (17), we find that M2σ → 6AN as N ≫ 1 and therefore the asymptotic

value behaves as δσk ∼ 1/√

− ln(−z/c).

We examine the scale-dependence of Pθ and Pσ in Fig. 7. The scale-invariance of Pθ is

preserved in the constant turn limit g = 0 where there is no phase transition and is mildly

broken with the increase of g. In the large g limit Pθ peaks around kc, which recovers the

finding in the hybrid inflation without a turning trajectory [57]. The scale-dependence of

Pσ is led by the heavy field approximation (36) in the IR regime.

16

Page 17: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

B. Bispectrum

We want to search for the imprint of the waterfall field in the three-point function of the

curvature perturbation ζ . For this purpose, we pick up interactions from field perturbations

at cubic order and compute their contribution to 〈ζ3〉 by using the in-in formalism [43–45].

The underlying assumption is that these higher-order corrections are perturbatively small

compared with the strongly coupled solutions obtained in the previous section. The most

important decay channel at tree-level shall come from the σ self-interaction [9, 18, 19]

L3 = −1

6a3R3Vσσσδσ

3 = −a3λR4σδσ3. (37)

This self-interaction is treated perturbatively provided that λRσ < H . There are some

other cubic interactions g2R2σδσδθδθ and g2R2θcδσδσδθ induced by the inflaton-isocurvaton

coupling. Having in mind that R ≫ Mp and g ∼ H , these interactions are less important

than (37) provided that λ > H2/R2.

Part of the conventional in-in formalism for the perturbative quasi-single field inflation [9]

is now adjusted for a strongly coupled system [18]. First of all, the definition (24) and (25)

in fact allows a direct decay of σ into θ (and vice versa), as can be seen by the non-vanishing

correlation function

〈δθk(z)δσp(z1)〉 = (2π)3δ3(k+ p)[

u+k (z)v

+∗p (z1) + u−

k (z)v−∗p (z1)

]

, (38)

where z1 ≡ k1η. It is also straightforward to define a mixed propagator

GR(k; z, z1) ≡ −iΘ(z − z1) [δθk(z), δσk(z1)] , (39)

= − i

(2π)3Θ(z − z1) 〈δθk(z)δσp(z1)〉 , (40)

where GR acts as the retarded Green’s function for both (28) and (29).

We then define the interaction picture field δθI , δσI as the solutions of the coupled

equations (28) and (29), which are evaluated by the EoM approach. To obtain the value

of correlation functions in the asymptotic limit −z → 0, it is more convenient to apply the

commutator representation of the in-in formalism [43–45]. At tree-level, the higher-order

corrections HI to the three-point function of δθ is then calculated by

〈δθ3(z)〉 =⟨(

∞∑

N=0

iN∫ z

dzN · · ·∫ z2

dz1[HI(z1), · · · [HI(zN), δθ3I (z)]]

)⟩

= i

∫ z

dz1⟨

0|[

HI(z1), δθ3I (z)

]

|0⟩

. (41)

17

Page 18: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

◆ g = 3.5

◆ g = 4

◆ g = 4.5

��-� ����� ����� ����� �

-���

-��

��

���

���θ/(�

��)

FIG. 8. The squeezed limit of the bispectrum Bθ as a function of c with g = {3.5, 4, 4.5} in the

unit of H. The dashed lines are the bestfit functions given by Eq. (45). The parameters are fixed

by λ = 10−6, θc = 1, p = 1 , R/Mp = 10 and H/Mp = 10−5.

By choosing HI(z) = λR4a4(z)σ(z)δσ3(z) from (37), where a(z) = −k/(Hz), the three-point

function (41) reads

〈δθ3(z)〉 = −2λR4 k31

H4Im

[∫ z dz1

z41σ(

z1k1

)GR(k1; z, z1)GR(k2; z,k2k1

z1)GR(k3; z,k3k1

z1)

]

. (42)

This expression can be identified as the mixed propagator approach of the Schwinger-Keldysh

diagrammatics interpretation from a set of decoupled initial states [21].

For convenience we shall fix k1 = k2 = kc and reparametrize k3 = ckc. The integration

(42) is now reduced as

〈δθ3(z)〉 = −2λR4 k3c

H4Im

[∫ z dz1

z41σ(

z1kc)G2

R(kc; z, z1)GR(ckc; z, cz1)

]

, (43)

where some of the numerical results are given in Fig. 8, and we have used the definition for

the bispectrum Bθ as

〈δθk1(z)δθk2

(z)δθk3(z)〉 ≡ (2π)4δ3(k1 + k2 + k3)Bθ(k1,k2,k3). (44)

Some of the numerical results of Bθ are given in Fig. 8. We have chosen the coupling

constant g > 3H so that the σ field can reach the new minimum in time for a given IR

18

Page 19: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

cutoff z = zf . The numerical results with g = 3.5H in Fig. 8 exhibit a similar shape as

the oscillatory component of the bispectrum from the constant turn inflation [20, 21]. It

is amplified in the direction towards the squeezed limit (c → 0). With the existence of a

waterfall phase transition, however, the oscillation component of Bθ becomes dominantly

large relative to the non-oscillatory contribution.

One can compare the results of Bθ with the bispectrum shape function Sθ used in [20, 21]

through the relation Bθ = (2π)3P 2θ Sθ/(k1k2k3)

2. Then the result in Fig. 8 is proportional

to c3Bθ ∝ c Sθ. In the equilateral limit where c → 1, the results can be nicely fitted by the

analytical form

c3

2λk3c

Bθ(c) = α1cβ1 cos (ω1 ln c− ϕ1) + α2c

β2 sin (ω2 ln c− ϕ2) + α3c2, (45)

based on the qualitative analysis constructed in [18, 19]. The term α3c2 comes from the de-

caying solution in δσ, which is negligible in the squeezed limit. The parameters of the oscilla-

tory terms depend on the mass m2σ. For a constant massm2

σ > 0 one shall find β1 = β2 = 3/2

and ω1 = ω2. Here we have the fitting results {β1, β2, ω1, ω2} = {−0.88,−0.88, 2.42, 2.49}for g = 3.5 and {β1, β2, ω1, ω2} = {−0.39,−0.44, 3.39,−2.49} for g = 4.

The results with g ≥ 4H are neither amplified in the squeezed limit nor in the equilateral

limit, and therefore they cannot be fitted by the analytical function (45). In this case a novel

shape of the oscillatory bispectrum is generated, with a maximal amplitude in between the

squeezed and the equilateral limits. This result is easy to be understood since the effective

mass of the isocurvaton perturbation (30) is changing from a large value at −z ≫ c to be

very small around −z = c and then becomes large again as z → zf . In the equilateral limit

where c < 0.01, the shape of the bispectrum with g ≥ 4H is the same as the cases with

a small coupling constant g. For a large coupling g, the mode function δσk with k ≪ kc

can decay to be much smaller than the asymptotic value δσIR given by (36), and it may

not reach back to δσIR in time. As a result, the contribution from modes δσk with k → 0

are suppressed in the integration (43), consistent with the finding from the constant turn

scenario with a large kinetic coupling case in [18, 19].

We have checked that the oscillation amplitude with g < 4H also starts to decay as we

extend the IR cutoff to c < 10−4. This means that the phase transition occurs early enough

such that the isocurvaton can find its new minimum in time even with a small g (see Fig. 2).

We thus conclude that, as long as −z/c is small enough, the result of c3Bθ always converges

19

Page 20: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

in the squeezed limit for arbitrary values of g/H > 0. Nevertheless for g2/H2 < 1 one can

treat δL2 given by (23) as perturbative interactions and proceed all the computation by the

canonical in-in formalism, as shown in [9].

We note that the shape of Bθ given by g = 3.5H could be what is really observed in this

scenario, provided that there is enough time for all the superhorizon modes to reach the

asymptotic value at the IR cutoff for observations.

VI. DISCUSSIONS

Particles with masses around the size of the Hubble scale during inflation are important

targets for the cosmological collider research. In this work we investigate the characteristic

signals from a class of scalar particles that experience of a waterfall phase transition accord-

ing to the mechanism introduced by the hybrid inflation [28]. As an explicit study case,

we incorporate the waterfall potential with the constant turn inflation [9], and we restrict

the parameter space to that of the quasi-single field regime where the energy density of the

isocurvaton is always subdominant.

With the help of the centrifugal force produced by the turning potential valley, the isocur-

vaton can obtain non-trivial initial conditions at the critical time for the phase transition,

depending on the model parameters. We numerically solved the background equation of

motion and found that the waterfall phase transition can last for more than 10 e-folds with

a sufficiently small coupling constant g = g/H . In the limit of g → 0, θc → ∞ and there is

no phase transition occur. Thus we are interested in the case with g ≥ H .

We have firstly checked the behavior of the long wavelength modes of the isocurvature

field under the waterfall phase transition. The long wavelength part of the waterfall field

is treated as an auxiliary classical field governed by the stochastic process [39, 40]. We

found that hybrid inflation in a turning potential valley can avoid the domain-wall formation

problem from a wide range of initial conditions. We also confirmed that the decomposition of

the waterfall field σ into a classical part σ0 and a quantum-fluctuation part δσ is generically

well-posed in this scenario.

We have resolved the quantum fluctuations δσ by virtue of equation of motion approach

[22] base on the modified in-in formalism for a strongly coupled system. We applied the

EoM method with mixed initial states to compute the two-point and three-point correlation

20

Page 21: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

functions. The appearance of a spiky feature in the power spectrum with a large g shows

that the time-translation invariance of the field perturbations in the IR limit is explicitly

broken by the waterfall phase transition.

Thanks to the waterfall phase transition, our numerical results show that the oscillatory

component in the bispectrum overwhelms the non-oscillatory contributions. One may obtain

two kinds of oscillating shape in this scenario. The first kind is a shape with an increasing

oscillating amplitude in the direction of the squeezed limit, corresponding to the case that the

IR cutoff for observations is taken at the time where all the isocurvaton field perturbations

can reach the asymptotic value δσIR given by (36). This shape is also demonstrated in

[20, 21] for the constant turn inflation, but with an additional magnification factor supplied.

The second kind is a new shape of bispectrum generated where the superhorizon modes in

the limit k → 0 are overdamped by the time of our observation. In this case the maximal

oscillating amplitude occurs in the intermediate region between the squeezed limit and the

equilateral limit. We shall remark that the increase of the oscillating amplitude from the

equilateral limit in both cases are generated even if the effective mass M2σ given by (30) is

always positive.

Finally, we want to emphasize that the purpose of the current study is to demonstrate

the existence of a novel oscillatory bispectrum which could be taken as the signature of a

waterfall phase transition during inflation. One may argue from the results of Fig. 7 that

the parameters with g/H > 3, used as our examples for the quantum clock signals in Fig.

8, are in fact ruled out by current observations if the critical scale kc locates in the CMB

region (namely 0.008Mpc−1 < kc < 0.1Mpc−1). A robust constraint of the model parameters

with observational data has been postponed as a future effort. It could be also interesting

to study the loop corrections in this scenario, although the EoM approach for quantum

fluctuations has effectively resummed all tree-level contributions from both perturbative

and non-perturbative interactions, given that loop corrections in a smooth phase transition

involved with tachyonic instability are found to be important if quantum fluctuations become

the dominant part of the isocurvature fields [58].

21

Page 22: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

ACKNOWLEDGMENTS

We thank Shi Pi, Teruaki Suyama and Rong-Fu Wu for several inspiring discussions, and

we appreciate the technical support from Xi Tong and Louis Yang. YW acknowledges the

International Symposium on Cosmology and Particle Astrophysics (CosPA) 2017 conference

held at Yukawa Institute for Theoretical Physics, Kyoto University (YITP-W-17-15), during

which part of the work was accomplished. YW is supported by ECS Grant 26300316 and

GRF Grant 16301917 from the Research Grants Council of Hong Kong. YPW is supported

by JSPS International Research Fellows and JSPS KAKENHI Grant-in-Aid for Scientific

Research No. 17F17322. JY was supported by JSPS KAKENHI Grant-in-Aid for Scientific

Research No. 15H02082, and Grant-in-Aid for Scientific Research on Innovative Areas No.

15H05888. SZ is supported by the the Hong Kong PhD Fellowship Scheme (HKPFS) issued

by the Research Grants Council (RGC) of Hong Kong.

[1] A. A. Starobinsky, Phys. Lett. 91B, 99 (1980); K. Sato, Mon. Not. Roy. Astron. Soc. 195,

467 (1981); A. H. Guth, Phys. Rev. D 23, 347 (1981); A. D. Linde, Phys. Lett. 108B, 389

(1982). A. Albrecht and P. J. Steinhardt, Phys. Rev. Lett. 48, 1220 (1982).

[2] K. Sato and J. Yokoyama, Int. J. Mod. Phys. D 24, no. 11, 1530025 (2015).

[3] C. Cheung, P. Creminelli, A. L. Fitzpatrick, J. Kaplan and L. Senatore, “The Effective Field

Theory of Inflation,” JHEP 0803, 014 (2008) [arXiv:0709.0293 [hep-th]].

[4] D. Baumann and L. McAllister, “Inflation and String Theory,” arXiv:1404.2601 [hep-th].

[5] J. M. Maldacena, “Non-Gaussian features of primordial fluctuations in single field inflationary

models,” JHEP 0305, 013 (2003) [astro-ph/0210603].

[6] P. Creminelli and M. Zaldarriaga, “Single field consistency relation for the 3-point function,”

JCAP 0410, 006 (2004) [astro-ph/0407059].

[7] N. Arkani-Hamed and J. Maldacena, “Cosmological Collider Physics,” arXiv:1503.08043 [hep-

th].

[8] X. Chen and Y. Wang, “Large non-Gaussianities with Intermediate Shapes from Quasi-Single

Field Inflation,” Phys. Rev. D 81, 063511 (2010) [arXiv:0909.0496 [astro-ph.CO]].

22

Page 23: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

[9] X. Chen and Y. Wang, “Quasi-Single Field Inflation and Non-Gaussianities,” JCAP 1004,

027 (2010) [arXiv:0911.3380 [hep-th]].

[10] E. Sefusatti, J. R. Fergusson, X. Chen and E. P. S. Shellard, “Effects and Detectability of

Quasi-Single Field Inflation in the Large-Scale Structure and Cosmic Microwave Background,”

JCAP 1208, 033 (2012). [arXiv:1204.6318 [astro-ph.CO]].

[11] J. Norena, L. Verde, G. Barenboim and C. Bosch, “Prospects for constraining the shape of

non-Gaussianity with the scale-dependent bias,” JCAP 1208, 019 (2012). [arXiv:1204.6324

[astro-ph.CO]].

[12] V. Assassi, D. Baumann and D. Green, “On Soft Limits of Inflationary Correlation Functions,”

JCAP 1211, 047 (2012) [arXiv:1204.4207 [hep-th]].

[13] E. Dimastrogiovanni, M. Fasiello and M. Kamionkowski, “Imprints of Massive Primordial

Fields on Large-Scale Structure,” JCAP 1602, 017 (2016). [arXiv:1504.05993 [astro-ph.CO]].

[14] F. Schmidt, N. E. Chisari and C. Dvorkin, “Imprint of inflation on galaxy shape correlations,”

JCAP 1510, no. 10, 032 (2015). [arXiv:1506.02671 [astro-ph.CO]].

[15] M. Yamaguchi and J. Yokoyama, Phys. Rev. D 74, 043523 (2006) [hep-ph/0512318].

[16] P. D. Meerburg, M. Mnchmeyer, J. B. Muoz and X. Chen, JCAP 1703, no. 03, 050 (2017)

[arXiv:1610.06559 [astro-ph.CO]].

[17] A. Moradinezhad Dizgah, H. Lee, J. B. Munoz and C. Dvorkin, “Galaxy Bispectrum from

Massive Spinning Particles,” arXiv:1801.07265 [astro-ph.CO].

[18] H. An, M. McAneny, A. K. Ridgway and M. B. Wise, “Quasi Single Field Inflation in the

non-perturbative regime,” arXiv:1706.09971 [hep-ph].

[19] H. An, M. McAneny, A. K. Ridgway and M. B. Wise, “Non-Gaussian Enhancements of Galac-

tic Halo Correlations in Quasi-Single Field Inflation,” arXiv:1711.02667 [hep-ph].

[20] X. Chen, M. H. Namjoo and Y. Wang, “Quantum Primordial Standard Clocks,” JCAP 1602,

no. 02, 013 (2016) [arXiv:1509.03930 [astro-ph.CO]].

[21] X. Chen, Y. Wang and Z. Z. Xianyu, “Schwinger-Keldysh Diagrammatics for Primordial

Perturbations,” JCAP 1712, no. 12, 006 (2017) [arXiv:1703.10166 [hep-th]].

[22] X. Chen, M. H. Namjoo and Y. Wang, “On the equation-of-motion versus in-in approach in

cosmological perturbation theory,” JCAP 1601, no. 01, 022 (2016) [arXiv:1505.03955 [astro-

ph.CO]].

23

Page 24: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

[23] X. Chen, “Primordial Features as Evidence for Inflation,” JCAP 1201, 038 (2012)

[arXiv:1104.1323 [hep-th]].

[24] R. Saito, M. Nakashima, Y. i. Takamizu and J. Yokoyama, “Resonant Signatures of Heavy

Scalar Fields in the Cosmic Microwave Background,” JCAP 1211, 036 (2012) [arXiv:1206.2164

[astro-ph.CO]].

[25] X. Chen, M. H. Namjoo and Y. Wang, “Models of the Primordial Standard Clock,” JCAP

1502, no. 02, 027 (2015) [arXiv:1411.2349 [astro-ph.CO]].

[26] R. Flauger, M. Mirbabayi, L. Senatore and E. Silverstein, “Productive Interactions: heavy

particles and non-Gaussianity,” JCAP 1710, no. 10, 058 (2017) [arXiv:1606.00513 [hep-th]].

[27] X. Tong, Y. Wang and S. Zhou, “Warm Quasi-Single Field Inflation,” arXiv:1801.05688 [hep-

th].

[28] A. D. Linde, “Hybrid inflation,” Phys. Rev. D 49, 748 (1994) [astro-ph/9307002].

[29] M. Li and Y. Wang, “Multi-Stream Inflation,” JCAP 0907, 033 (2009) [arXiv:0903.2123

[hep-th]].

[30] N. Afshordi, A. Slosar and Y. Wang, “A Theory of a Spot,” JCAP 1101, 019 (2011)

[arXiv:1006.5021 [astro-ph.CO]].

[31] J. Liu, Y. Wang and S. Zhou, “Nonuniqueness of classical inflationary trajectories on a high-

dimensional landscape,” Phys. Rev. D 91, no. 10, 103525 (2015) [arXiv:1501.06785 [hep-th]].

[32] J. Yokoyama, Phys. Rev. Lett. 63, 712 (1989); J. Yokoyama, Phys. Lett. B 212, 273 (1988);

J. Yokoyama, Phys. Lett. B 231, 49 (1989).

[33] M. Nagasawa and J. Yokoyama, “Phase transitions triggered by quantum fluctuations in the

inflationary universe,” Nucl. Phys. B 370, 472 (1992).

[34] S. Clesse, “Hybrid inflation along waterfall trajectories,” Phys. Rev. D 83, 063518 (2011)

[arXiv:1006.4522 [gr-qc]].

[35] J. Garcia-Bellido, A. D. Linde and D. Wands, “Density perturbations and black hole formation

in hybrid inflation,” Phys. Rev. D 54, 6040 (1996) [astro-ph/9605094].

[36] J. Silk and M. S. Turner, Phys. Rev. D 35, 419 (1987); D. Polarski and A. A. Starobinsky,

Nucl. Phys. B 385, 623 (1992). .

[37] M. Yamaguchi and J. Yokoyama, “Chaotic hybrid new inflation in supergravity with a run-

ning spectral index,” Phys. Rev. D 68, 123520 (2003) [hep-ph/0307373]; M. Yamaguchi and

J. Yokoyama, “Smooth hybrid inflation in supergravity with a running spectral index and

24

Page 25: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

early star formation,” Phys. Rev. D 70, 023513 (2004) [hep-ph/0402282].

[38] M. Kawasaki, T. Takayama, M. Yamaguchi and J. Yokoyama, “Power Spectrum of the Density

Perturbations From Smooth Hybrid New Inflation Model,” Phys. Rev. D 74, 043525 (2006)

[hep-ph/0605271].

[39] A.A. Starobinsky, Stochastic de sitter (inflationary) stage in the early universe, in Current

topics in field theory, quantum gravity and strings, H.J. de Vega and N. Sanchez eds., Springer

(1986).

[40] A. A. Starobinsky and J. Yokoyama, Phys. Rev. D 50, 6357 (1994) [astro-ph/9407016].

[41] J. O. Gong and M. Sasaki, “Waterfall field in hybrid inflation and curvature perturbation,”

JCAP 1103, 028 (2011) [arXiv:1010.3405 [astro-ph.CO]].

[42] D. S. Salopek, J. R. Bond and J. M. Bardeen, Phys. Rev. D 40, 1753 (1989).

[43] S. Weinberg, “Quantum contributions to cosmological correlations,” Phys. Rev. D 72, 043514

(2005) [hep-th/0506236].

[44] X. Chen, “Primordial Non-Gaussianities from Inflation Models,” Adv. Astron. 2010, 638979

(2010) [arXiv:1002.1416 [astro-ph.CO]].

[45] Y. Wang, “Inflation, Cosmic Perturbations and Non-Gaussianities,” Commun. Theor. Phys.

62, 109 (2014) [arXiv:1303.1523 [hep-th]].

[46] A. J. Tolley and M. Wyman, “The Gelaton Scenario: Equilateral non-Gaussianity from multi-

field dynamics,” Phys. Rev. D 81, 043502 (2010) [arXiv:0910.1853 [hep-th]].

[47] A. Achucarro, J. O. Gong, S. Hardeman, G. A. Palma and S. P. Patil, “Mass hierar-

chies and non-decoupling in multi-scalar field dynamics,” Phys. Rev. D 84, 043502 (2011)

[arXiv:1005.3848 [hep-th]].

[48] D. Baumann and D. Green, “Equilateral Non-Gaussianity and New Physics on the Horizon,”

JCAP 1109, 014 (2011) [arXiv:1102.5343 [hep-th]].

[49] A. Achucarro, J. O. Gong, S. Hardeman, G. A. Palma and S. P. Patil, “Effective theories

of single field inflation when heavy fields matter,” JHEP 1205, 066 (2012) [arXiv:1201.6342

[hep-th]].

[50] X. Chen and Y. Wang, “Quasi-Single Field Inflation with Large Mass,” JCAP 1209 (2012)

021 [arXiv:1205.0160 [hep-th]].

[51] S. Pi and M. Sasaki, “Curvature Perturbation Spectrum in Two-field Inflation with a Turning

Trajectory,” JCAP 1210, 051 (2012) [arXiv:1205.0161 [hep-th]].

25

Page 26: Jockey Club Institute for Advanced Study, arXiv:1804.07541v1 … · 2018-04-23 · phase transition in the isocurvature direction driven by the symmetry-breaking mechanism based on

[52] A. Achucarro, V. Atal, S. Cespedes, J. O. Gong, G. A. Palma and S. P. Patil, “Heavy fields,

reduced speeds of sound and decoupling during inflation,” Phys. Rev. D 86, 121301 (2012)

[arXiv:1205.0710 [hep-th]].

[53] R. Gwyn, G. A. Palma, M. Sakellariadou and S. Sypsas, “Effective field theory of weakly

coupled inflationary models,” JCAP 1304, 004 (2013) [arXiv:1210.3020 [hep-th]].

[54] J. O. Gong, S. Pi and M. Sasaki, “Equilateral non-Gaussianity from heavy fields,” JCAP

1311 (2013) 043 [arXiv:1306.3691 [hep-th]].

[55] X. Tong, Y. Wang and S. Zhou, JCAP 1711, no. 11, 045 (2017) [arXiv:1708.01709 [astro-

ph.CO]].

[56] A. V. Iyer, S. Pi, Y. Wang, Z. Wang and S. Zhou, JCAP 1801, no. 01, 041 (2018)

[arXiv:1710.03054 [hep-th]].

[57] A. A. Abolhasani, H. Firouzjahi, S. Khosravi and M. Sasaki, “Local Features with Large Spiky

non-Gaussianities during Inflation,” JCAP 1211, 012 (2012) [arXiv:1204.3722 [astro-ph.CO]].

[58] Y. P. Wu and J. Yokoyama, arXiv:1704.05026 [hep-th].

26