I Introduction
Having a large blue tilt in the axionic isocurvature spectrum allows
cold dark matter (CDM) density perturbations to be enhanced on short
length scales without being in conflict with the precision cosmology
that exists for scales Mpc-1 (Chluba:2013dna, ; Takeuchi2014, ; Dent:2012ne, ; Chung:2015pga, ; Chung:2015tha, ; Chluba:2016bvg, ; Chung:2017uzc, ; Planck:2018jri, ; Chabanier:2019eai, ; Lee:2021bmn, ; Kurmus:2022guy, ).
The generation of isocurvature perturbations by spectator axions,
its model-specific characteristics, and the related observational
limitations have been extensively investigated in the past (see for
example (Kasuya1997, ; Kawasaki1995, ; Nakayama2015, ; Harigaya2015, ; Kadota2014, ; Kitajima2014, ; Kawasaki2014, ; Higaki2014, ; Jeong2013, ; Kobayashi2013, ; Hamann2009, ; Hertzberg2008, ; Beltran2006, ; Fox2004, ; Estevez2016, ; Kearney2016, ; Nomura2015, ; Kadota2015, ; Hikage:2012be, ; Langlois2003, ; Mollerach1990, ; Axenides1983, ; Jo2020, ; Iso2021, ; Bae2018, ; Visinelli2017, ; Takeuchi:2013hza, ; Bucher:2000hy, ; Lu:2021gso, ; Sakharov:2021dim, ; Rosa:2021gbe, ; Jukko:2021hql, ; Chen:2021wcf, ; Jeong:2022kdr, ; Cicoli:2022fzy, ; Koutsangelas:2022lte, ; Kawasaki:2023zpd, )).
Although there are models of axion which naturally generate large
blue tilted spectra when there are no quartic potential terms in the
radial field (Kasuya:2009up, ; Dreiner:2014eda, ), there is no previous
discussion in the literature regarding generating a large range
of very blue spectrum followed by a plateau from a well-motivated
axion models that contain a quartic term in the radial potential (Kim:2008hd, ; DiLuzio:2020wdo, ). From a model building perspective, one can therefore ask whether
the overdamped spectrum of (Kasuya:2009up, ) (i.e. a smooth spectrum
composed of an exponentially large -range with a very blue spectral
index and followed by a zero spectral index plateau without any large
bumplike features) can be a signature of flat direction models that
are distinct from the quartic radial potential models. If the answer
is affirmative, then not only is the time-dependent mass a property
that one can infer (Chung:2015tha, ) from measuring the spectral
shape similar to that of (Kasuya:2009up, ), also the existence
of flat direction would be inferrable from such a measurement.
Motivated by this question and also from the desire to find novel
well-motivated beyond the Standard Model scenarios that generate a
strongly blue tilted axionic isocurvature spectra, we consider a generic
complex scalar sector with the radial direction field , the
angular field , and a quartic coupling . The quartic
term usually controls the axion decay constant
|
|
|
(1) |
(what people often denote as ) where the mass parameter
controls the tachyonic mass term responsible for spontaneous
breaking of Peccei-Quinn (PQ) symmetry. For the blue isocurvature
models, we require a large rolling period of the radial field
because it is the time-dependence of the background fields that map
to the nontrivial positive power (i.e. blue tilt) of in the
dimensionless power spectrum. When
and the initial kinetic energy is negligible, we might naively expect
to roll to on a time scale of .
Such a fast roll for (making a large blue
tilt) would generate the range of over which the blue spectra
is produced to be
|
|
|
(2) |
where is the length scale that leaves the
horizon when begins to roll and is
the scale that leaves the horizon when reaches the minimum
such that modes having will approximately
be a flat spectrum.
However, if there is an axion background field motion in the conserved
angular direction, we know that the time scale
to reach can be infinite in the limit that
all symmetry breaking terms are turned off and
. Hence, in scenarios where the angular motion
in the conserved angular momentum direction is large, we might expect
to be able to have a similar radial rolling as the flat direction
model of (Kasuya:2009up, ). Unlike in the scenarios of (Co:2019wyp, ; Co:2021lkc, ),
we will use the initial conditions where the phenomenology generating
rotations are occurring during inflation. In such angular momentum
dependent scenarios, one naively expects the main limitations to obtaining
a large to be the dilution
of the angular momentum due to the Hubble expansion. After substituting
the kinetic derived for the “angular momentum”
, there is the well known effective potential term
|
|
|
(3) |
which naively indicates that the effect of will decay as
, becoming irrelevant too fast to be of interest. However,
the situation is a bit more interesting.
Because is decreasing when ,
the denominator initially decreases only like :
i.e. more mildly, giving intuitively a better chance for a blue spectrum
to be generated for a larger number of efolds. Furthermore, this means
that is also decreasing as , making the relative
contribution of the angular momentum not diminish with the increasing
scale factor. Indeed, this causes the potential to scale as the inverse
mass dimension of the potential, hinting that this is a conformal
limit. As will be explained in this paper, this is a time-independent
conformal limit of a special type (different from a massless scalar
field in Minkowski space or a massless equilibrium axion in dS space),
and this will be utilized to generate a blue isocurvature spectrum
for the axion field which is approximately : i.e.
corresponding to a spectral index of .
In addition to constructing a novel model of generating a blue spectrum,
we also systematically quantize the fields in this background-out-of-equilibrium
situation which can be characterized by the novel nonvanishing of
the commutator
representing velocity correlation even in the absence of non-derivative
correlations. Although the work of (Creminelli:2023kze, ; Hui:2023pxc, )
quantizes a similar theory and agrees with our results, we present some distinct and unique
details here regarding the axion spectrum (particularly regarding
conformal symmetry representation) and apply it to isocurvature and
dark matter phenomenology. One revelation is that the radial perturbations
about the conformal background solution mix with the angular perturbations
for any eigenstate of the Hamiltonian even at the quadratic fluctuation
level, and the different energy eigenvectors (each eigenvector representing
the mixing) are not orthogonal. Indeed, it will be shown that a massive
that kinetically mixes with has nearly
identical conformal representation as . A more important
revelation is that, despite the complicated quantum mode mixing arising
from the time-dependent background, explicit quantization allows one
to construct a time-independent Hamiltonian whose ground state well-represents
the vacuum. Because of this and angular field translational symmetry,
Goldstone theorem still applies during the conformal period, and the
dispersion relationship is approximately linear in as
but with a different sound speed coefficient of , similar
to a relativistic perfect fluid pressure wave. Indeed, it is well
known that a quartic complex scalar with spontaneous breaking
is a simple model of a superfluid (see e.g. (Leggett:1999zz, )).
As far as model parameters are concerned, there are the initial conditions
of the background fields, the quartic coupling, and the usual axion
parameters which control the dark matter abundances. The main theoretical
limitation on extending this blue spectrum over a large range
is the requirement that the axion remains a spectator, which limits
the coupling and the background field initial displacement value in
the conformal regime. We also identify a range of initial condition
deformations away from the conformal limit over which the isocurvature
spectrum is approximate , beyond which parametric resonance
sets in and destroys the smooth blue spectrum. We identify the parameter
regime in which this type of model can reproduce a spectrum of blue-tilt
followed by a plateau.
The order of presentation will be as follows. In Sec. II,
we define the notation for the “vanilla” axion model and make
general arguments of how a time-independent conformal limit and the
spectral index arises with the combination of large field
displacements and angular momentum. In Sec. III,
we quantize the theory explicitly about the large phase angular momentum
to make the vacuum choice precise and to compute the resulting normalization
for the desired correlation function. We also give a simplified discussion
of how the intermediate-time transition away from the time-independent
conformal-era will not result in a large bump in the isocurvature
spectrum. In Sec. IV, we discuss how
deformations of the initial conditions away from the time-independent
conformal limit will modify the spectrum. This will lead to oscillatory
features in the spectrum. In Sec. V, we present
example isocurvature spectra plots and the parametric ranges over
which the QCD axion phenomenology is compatible with observations.
We then conclude with a summary. Many appendices follow that provide
details of the results presented in the main body of the work. For
example, the details of the conformal field representation will be
given in Appendix A and the details of
the quantization is presented in Appendix D.
VI Conclusion
In this paper, we have shown that modifying the initial conditions
of a generic symmetric quartic potential complex scalar model
can lead to a novel axion isocurvature scenario in which a transition
takes place from a time-independent conformal phase to the time-dependent
conformal phase, the latter being the usual equilibrium axion scenario.
Such time-independent spontaneously broken conformal phase initial
condition is controlled by a large classical background phase angular
momentum
and a large radial field displacement .
With such initial conditions for the background, the quantum perturbations
remarkably enter a nontrivial time-independent spontaneously symmetry-broken
conformal phase characterized by a long wavelength spectral index
of and a Goldstone dispersion with a sound speed of .
Interestingly, the cross-correlation function
during this time-independent conformal phase between the radial and
the axion fields does not vanish even though
to leading order in perturbation theory.
After the reaches the usual spontaneous PQ symmetry
breaking minimum, the theory enters the usual time-dependent conformal
phase characterized by the time-dependent effective mass term .
For values corresponding to this time region, denoted as ,
the isocurvature spectrum is the well-known flat plateau,
and the Goldstone dispersion has a sound speed of . One nontrivial
phenomenological result established in this paper is that the spectral
transition from to for realistic parameter ranges
can be sudden such that there is no large bump connecting these two
regions. This means that this quartic potential model can behave qualitatively
differently from the overdamped supersymmetric (SUSY) scenarios of
(Kasuya:2009up, ) where there is a bump (Chung:2016wvv, ).
Furthermore, if the range over which the blue spectral index
sets in is sufficiently large, then the present model becomes more
fine tuned compared to the flat direction models. In the sense of
making parameters less tuned, the SUSY models in this context can
be considered analogous to the low-energy SUSY models solving the
Higgs mass hierarchy problem.
With two-parameter initial condition perturbations away from those
generating the time-independent spontaneously broken conformal phase,
we have shown that the smooth to spectra transition
scenarios are stable with deformations of
and .
On the other hand, small deviations of the spectral amplitudes linear
in these deformation ratios eventually gain a nonlinear dependence
as these deviations grow beyond magnitudes of around . Afterwards
parametric resonances strongly set in and destroy the original qualitative
shape of the spectra. The small oscillatory features apparent in small
deformation cases are well-fit by a simple formula characteristic
of the sound speed of the conformal phase.
We have also explored the parametric region for which this scenario
is phenomenologically interesting. Requiring the simultaneous
satisfaction of constraint of the axion being a QCD axion, maximum
blue spectral interval satisfying
, quartic coupling of order unity,
all of the DM being composed of axions, and isocurvature not violating
the current bounds, no viable parameter region exists. On the other
hand, with the relaxation of these constraints, there is a phenomenologically
viable parametric region as shown in Fig. 13.
Because the energy density rises steeply compared to the flat direction
scenarios as the radial field is displaced, the spectator condition
imposes a significant constraint that makes this scenario sensitive
to the quartic coupling.
There are many natural future directions to explore. Given the natural
similarities between this model and the SUSY flat direction model
of (Kasuya:2009up, ), it would be interesting to see whether
non-Gaussianities can break the degeneracy. Indeed, there is a peculiar
feature of the time-independent conformal spectra which kinetically
cross correlates the radial mode and the axial mode, and this kinetic
mixing does not exist in the SUSY flat direction model. Hence, we
would expect the non-Gaussianities to be different between the two
models even if the isocurvature spectra are similar. Another interesting
direction is in exploring the observability of the oscillatory features
in the power spectra. As noted above, in the quasi-conformal model,
there are oscillatory features in the isocurvature spectra for small
deviations away from time-independent conformality and since those
oscillations encode the sound speed information, it
would be interesting to see if observations can measure this sound
speed. Of course, work even remains to be done in assessing the observability
of the oscillatory features in the underdamped SUSY models (Chung:2021lfg, )
as noted in (Chung:2023syw, ).
Appendix A Conformal limit for the background
In this section, we describe how a large and
limit together with a certain classical boundary condition corresponds
to a spontaneously broken approximate conformal limit of the field
theory of and during which
where is the scale factor corresponding to the metric
|
|
|
(191) |
We begin by deriving the effective action from a general symmetric
renormalizable theory that spontaneously breaks an approximate conformal
symmetry with a large phase angular momentum. We then use the conformal
symmetry parameterization to generate an automorphism of the correlation
functions. This allows one to derive a differential equation for the
correlation functions whose general solution is given. We will then
use the spontaneously broken coset representation to derive
for the correlators and use the
absence of this symmetry for correlators to argue for
the
dependence.
Start with a general renormalizable invariant action action
given by Eq. (16):
|
|
|
|
(192) |
where . Note that this theory is almost invariant
under the following constant scaling conformal (dilatation) transform:
|
|
|
(193) |
were it not for the term. Look for constant
solutions to the equation of motion for :
|
|
|
(194) |
where we used the generated conservation law to set
|
|
|
(195) |
with being a constant. Because metric scaling will be involved
later, here we have chosen to keep explicit coming from
the conserved quantity being proportional to the charge density
and not its associated 1-form . Eq. (194)
has an approximately time-independent solution
|
|
|
(196) |
when
|
|
|
(197) |
Substituting Eq. (195) into (196),
we find
|
|
|
(198) |
which is a constant by the virtue of Eq. (196). Since
is a constant, we know from this equation that
is a constant. In terms of and fields, these solutions
represent
|
|
|
(199) |
|
|
|
(200) |
indicating that this is a nontrivial approximate time-dependent background
in terms of the canonical real radial field.
Now, define
|
|
|
(201) |
By neglecting and consistently with Eq. (197),
the action now turns out to be completely independent of the scale
factor :
|
|
|
(202) |
showing explicitly that we have a conformal theory enjoying the symmetry
|
|
|
|
(203) |
where is a constant.
Here, the arguments and of are viewed
as externally input parameters, and we transform them as we would
a spurion. Note that the conformal representation here is different
from the dilatation subgroup representation of diffeomorphism (see
e.g. (Ginsparg:1988ui, )) especially because we are scaling
which is a parameter, as in a spurion representation of the
conformal group in a free massive scalar theory. On the other hand,
rewriting Eq. (199) as
|
|
|
(204) |
shows that the transform
comes from scaling the scale factor by a constant as ,
such that the symmetry of Eq. (203) being an element
of the conformal group is evident. We will see how these symmetries together with diffeomorphism will
give rise to constraints on the correlation functions of interest
below.
Expand the fields as
|
|
|
(205) |
|
|
|
(206) |
where is given by Eq. (200)
and look for the effective action governing the perturbations only.
The perturbation-only action is
|
|
|
|
(207) |
|
|
|
|
(208) |
which enjoys the conformal symmetry
|
|
|
(209) |
where the constant does not transform.
However, as we will see below, will transform
under diffeomorphism because of the time derivative.
Carry out a coordinate change (diffeomorphism)
leading to
|
|
|
|
(210) |
and
leading to the diffeomorphism invariant action transforming as
|
|
|
|
(211) |
|
|
|
|
(212) |
Scaling variables, we find
|
|
|
|
(213) |
Because of Eq. (209), this is equivalent
to
|
|
|
(214) |
and thus
|
|
|
|
(215) |
|
|
|
|
(216) |
Let’s see the implication of this on the Feynman correlator which
will be equivalent to the in-in equal time correlator that we seek
at free field level:
|
|
|
(217) |
Change variables
|
|
|
(218) |
to conclude
|
|
|
|
(219) |
or more explicitly
|
|
|
(220) |
To derive the differential equation governing the correlator by assuming
and spatial translation invariance, start by writing
|
|
|
(221) |
where is a function of variables .
This and Eq. (220) says
|
|
|
(222) |
Taking a derivative with respect to and setting , we find
|
|
|
|
|
|
|
(223) |
governing the correlator. A general solution to this equation is
|
|
|
(224) |
where is a general function of two variables.
Now, time translation invariance implies being time-translation-invariant.
This results in the differential equation
|
|
|
(225) |
whose general solution is
|
|
|
(226) |
giving
|
|
|
|
(227) |
Furthermore, we know for circular orbits
|
|
|
(228) |
and thus combine and dependence
to conclude
|
|
|
(229) |
|
|
|
(230) |
which are manifestly time-independent but contain arbitrary
dependences unlike in the case of massless scalar fields in Minkowski
space.
Next, note the induced shift symmetry
has an associated current
|
|
|
(231) |
whose conservation is
|
|
|
(232) |
which is notably linear. In normal mode-Fourier space, Eq. (232)
is
|
|
|
(233) |
In the large limit, we have the usual Goldstone condition
for any nonvanishing . The fact that the two
modes have dispersion possibilities can be viewed
as a consequence of approximately in tact CPT symmetry in that limit.
For small , there is at least one massless mode that is independent
of and as long as
and do not diverge. Let’s label that massless
mode frequency as . Indeed, because of the
dependent mixing in Eq. (208), both
and do not
vanish.
According to Eq. (208), we see that the Lagrangian has
a mode contribution
|
|
|
(234) |
|
|
|
(240) |
which in the
limit has decoupling from and
not changing the approximate correlation function for .
In that approximation, the factor
in Eq. (229) has an expansion for large
as
|
|
|
(241) |
where is independent of and
is a function where . This and Eq. (229)
implies the correlator in Fourier space behaving as corresponding
to a spectral index of matching Eq. (28).
Hence, unlike the ordinary massless Minkowski field, one has to use
Goldstone dynamical information contained in Eq. (240)
to fix the scaling for the
correlator.
With this same
approximation, we see from Eq. (240) that a decoupled
becomes infinitely heavy as .
Since this means that
should vanish with the same limit, we expect
|
|
|
(242) |
if the expansion is analytic in inverse powers of .
Explicit computations justify the analyticity.
Now, the sound speed of cannot quite be read off from
this expression since
generates a mixing between and .
This mixing generated change in the sound speed, which is the most
theoretically interesting aspect of the system studied in this paper,
and other aspects of this system are addressed in the main body of
the text when we quantize the theory.
Appendix B WKB approximation for oscillating potentials
In this section, we explain Eq. (261) which is a
generalization of the WKB ansatz applicable for dispersion relationships
with a fast time-oscillation component.
Consider the following differential equation
|
|
|
(243) |
The WKB method allows us to approximate the solution to the above
differential equation as
|
|
|
(244) |
given
|
|
|
(245) |
and hence the mass-squared function must be slowly varying.
Let us now consider the situation where the mass-squared term is
|
|
|
(246) |
where are constants and we are interested in cases with
and such that the mass-squared function is characterized
by high frequency large amplitude oscillations. It is easy to note
that the WKB approximation as given above is inappropriate and diverges
at the zeros of . Although one may use matching solutions
at the zero crossings, such an approach is unwieldy for a fast oscillating
potential and doesn’t capture the long-time characteristic behavior
of the system.
Another well-known approach begins with noting that Eq. (243)
with the mass-squared function given in Eq. (246) satisfies
the Mathieu differential equation
|
|
|
(247) |
where
|
|
|
|
(248) |
|
|
|
|
(249) |
|
|
|
|
(250) |
with the generalized solution
|
|
|
(251) |
|
|
|
(252) |
for , as the generalized angular Mathieu functions. In
the limiting case , the Mathieu function has the following
series expansion in powers of
|
|
|
(253) |
where
|
|
|
(254) |
Using this, we can identify
|
|
|
(255) |
in the limit
|
|
|
(256) |
and the exact Mathieu solution given in Eq. (251)
can be approximated up to first order in as
|
|
|
(257) |
Hence, in the limit , we observe that the solution
to the oscillatory mass-squared function in Eq. (246)
is a superposition of states with the dominant state having a frequency
. Therefore, up to an accuracy of , the WKB approximate
solution to the differential Eq. (243) can be
given as
|
|
|
(258) |
where
|
|
|
(259) |
Note that as long as , the dominant frequency of the WKB
approximation is given by the slow-varying mass parameter such that
the WKB method no longer suffers from any oscillatory divergences.
The above results motivate us to draw following important conclusions.
Given a differential equation of the form
|
|
|
(260) |
where and are slow-varying non-oscillatory functions,
the solution to the above differential equation can be approximate
through the following WKB ansatz
|
|
|
(261) |
if
|
|
|
(262) |
where
|
|
|
(263) |
For , then the solution exhibits a large hierarchy in states
such that the system can be described as a superposition of IR and
UV states
|
|
|
(264) |
with the corresponding frequencies as
|
|
|
|
(265) |
|
|
|
|
(266) |
and the amplitudes
|
|
|
(267) |
Appendix D Details of quantization
Here we present the details of the quantization of the radial-axion
system in the presence of large phase angular momentum background
classical solution. The nontriviality will be coming from the nonvanishing
of the cross-commutator
even though . This quantization is what allows
us to compute the sound speed and the vacuum structure rigorously.
In terms of the linear order field fluctuations
where , we derive the EoM from
the action in Eq. (11) using the Euler-Lagrange equation
|
|
|
(278) |
where is the component of the Lagrangian which
is quadratic order in linear perturbations. The EoM for
and are obtained from the above expression as
|
|
|
(279) |
|
|
|
(280) |
Eqs. (279) and (280) form a system of coupled
ODEs and can be expressed compactly as
|
|
|
|
(281) |
where
|
|
|
(282) |
and
|
|
|
(283) |
Note that the linear perturbations are kinetically
coupled through the coefficient . We will
refer to all scenarios where
as strongly coupled. By defining new field variables ,
we can rewrite the above system of equations as
|
|
|
|
(284) |
where we modify as
|
|
|
|
(287) |
and .
We will quantize this system of coupled fields
using the commutator relations defined in Eq. (35).
From the Lagrangian, we find the canonical momenta as
|
|
|
(288) |
and
|
|
|
|
(289) |
Hence, we arrive at the following commutator expressions for the fields
and their time-derivatives :
|
|
|
|
|
|
|
|
|
|
|
|
(292) |
which is remarkable since while .
As expected in the decoupling limit when ,
the kinetic cross commutators vanish.
Next we write the most general solution for in terms
of time-independent non-Hermitian ladder operators and
mode function as
|
|
|
(293) |
where is the flavor index and counts the number of distinct
frequency solutions. The time-derivative of the field is
|
|
|
(294) |
The ladder operators satisfy relation
|
|
|
(295) |
where the coefficients must be determined by solving for
mode function and using canonical commutator relation
given in Eq. (292). Substituting our general solution
from Eq. (293) into Eq. (284) we obtain
|
|
|
|
(296) |
where
|
|
|
(297) |
|
|
|
|
(300) |
D.1 Normal modes
We will now solve the system of equations given in Eq. (296)
during the conformal regime. Hence, we propose the following ansatz
|
|
|
|
(305) |
for time-independent mode vectors .
Substituting this ansatz into Eq. (296) we obtain
|
|
|
|
(308) |
which is rewritten as
|
|
|
(309) |
where we have replaced with the conserved
angular momentum from Eq. (12) and .
Since we solve Eq. (296) at an early time
and in the conformal limit
and , we can neglect any small
amplitude oscillations of the background radial field. In the conformal
limit ,
and hence we arrive at the reduced expression for Eq. (309)
given as
|
|
|
(310) |
We note that in the conformal limit, our system is defined by time-independent
coefficients. The distinct “real” frequency solutions obtained
by solving Eq. (310) are
|
|
|
(311) |
where
|
|
|
(312) |
and . In the IR limit
|
|
|
(313) |
the two distinct frequency squared
|
|
|
(314) |
and
|
|
|
(315) |
correspond to low and high frequency solutions and are separated by
the large hierarchy. In the
UV limit,
|
|
|
(316) |
and the two frequency solutions become degenerate.
We write the full mode function as
|
|
|
|
(323) |
|
|
|
|
(328) |
where counts distinct frequencies
given by Eq. (312). The normal vectors corresponding
to each frequency are given by the ratios
|
|
|
|
(329) |
|
|
|
|
(330) |
which are purely “imaginary”. Hence, we can rewrite the solution
for the mode function as
|
|
|
|
(337) |
|
|
|
|
(342) |
where the frequencies are constant, real and positive
and the ratios are purely imaginary. In the
decoupling limit where the kinetic terms mixing
and vanish and the decoupled solution becomes :
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where the decoupled “instantaneous” frequencies are
|
|
|
(343) |
for
|
|
|
|
(344) |
|
|
|
|
(345) |
Substituting our solution for from Eq. (342)
into the expression for we obtain
|
|
|
|
|
|
|
|
(346) |
where
|
|
|
(347) |
and
|
|
|
(348) |
D.2 Ladder algebra
To evaluate the ladder algebra, we first express the ladder operators:
, and their conjugates in terms
of the fields and its conjugate momenta .
To this end, we define
|
|
|
(349) |
and
|
|
|
(350) |
From the above definition and Eq. (346),
we conclude that
|
|
|
|
(351) |
|
|
|
|
|
|
|
|
(352) |
where is the flavor index of our two-field coupled system and
runs from to . From Eqs. (351) and (352)
we setup the following system of equations to solve for the ladder
operators
|
|
|
(353) |
Solving the above equation yields
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(354) |
where we use Eq. (347) to set
|
|
|
(355) |
We can summarize these results as
|
|
|
|
(356) |
|
|
|
|
(357) |
where . The and operators appearing
on the RHS of Eq. (353) can be evaluated
as
|
|
|
|
(358) |
|
|
|
|
(359) |
|
|
|
|
(360) |
and,
|
|
|
|
(361) |
|
|
|
|
(362) |
|
|
|
|
(363) |
where we used .
It follows then that the commutators of and operators
can be obtained from the commutator relations in Eq. (292).
Hence,
|
|
|
|
|
|
|
|
|
|
|
|
(364) |
|
|
|
|
|
|
|
|
|
|
|
|
D.3 Ladder commutators
Using the solution for the ladder operators
and their conjugates as given in Eq. (354), and the
commutator algebra of operators given in Eq. (364),
we can evaluate the ladder commutators. There are 6 unique combinations
of the commutators for frequencies that we present below.
We work out the first commutator in detail and leave the remaining
for the readers as an exercise:
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(365) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Hence for the operator set defined as
|
|
|
(366) |
the commutators are
|
|
|
(367) |
where we set the coefficients through the expressions
|
|
|
|
(368) |
|
|
|
|
(369) |
Clearly, the ladder operators associated with distinct frequencies
commute with each other. Hence, we can define a common vacuum state
which is annihilated simultaneously by both
and for all :
|
|
|
|
(370) |
|
|
|
|
(371) |
Note that the normal mode vectors associated with
the corresponding ladder operators are not orthogonal.
D.4 Hamiltonian
We conclude our discussion on the quantization of the coupled system
by evaluating its Hamiltonian. We show that in the conformal limit,
the normal mode solution given in Eq. (342) diagonalizes
the Hamiltonian such that the vacuum state
is the state of minimum energy.
Hence, let us consider the Hamiltonian density defined through the
expression
|
|
|
(372) |
and the Hamiltonian given by
|
|
|
(373) |
In the time-independent conformal regime when and
are constants, the Hamiltonian simplifies to
|
|
|
|
|
|
|
|
(374) |
We introduce Fourier notation
|
|
|
|
(375) |
|
|
|
|
(376) |
|
|
|
|
(377) |
|
|
|
|
(378) |
Taking the Fourier transform of the fields we write the Hamiltonian
as
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where
|
|
|
|
(381) |
|
|
|
|
(384) |
Hence, in the conformal limit the Hamiltonian is diagonal in terms
of the fields and its time-derivatives.
Using our general solution for the fields
|
|
|
(385) |
it is possible to write
|
|
|
(386) |
where the indices . Below we evaluate :
|
|
|
|
|
|
|
|
(387) |
Using the mode solution
|
|
|
(388) |
it follows that
|
|
|
|
|
|
|
|
In the above equation, the expression in the second bracket goes to
when . Hence, we conclude that for all combinations
of . Similar calculations show that .
Next, we evaluate :
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
which is non zero only when . Similarly
|
|
|
|
|
|
|
|
vanishes when . Hence we find that for the amplitudes ,
the coefficients vanish for all combinations of .
Meanwhile, we find that are non-zero only when .
Thus, the normal frequency solutions corresponding to
and diagonalize our Hamiltonian, which we write as
|
|
|
|
|
|
|
|
|
|
|
|
Using the expression derived for the coefficients from
Eq. (368) we find that
|
|
|
(389) |
which allows us to write the final form of the Hamiltonian as
|
|
|
|
|
|
|
|
The vacuum state when applied to the above
Hamiltonian results in the lowest energy state with ground state energy
. Also,
we see that the one particle state
is an eigenstate of the Hamiltonian with the energy eigenvalue .
We note that any other choice for the mode amplitudes
other than that given in Eqs. (329) and (330)
will lead to a higher energy for the vacuum state
and as such it would not be the correct ground state of our theory.
Appendix F Relationship between radial
and angular modes
Suppose we parameterize a sigma model with the symmetry
as
|
|
|
(403) |
There is a shift symmetry
|
|
|
(404) |
where is a constant. If obeys a linear equation
of motion
|
|
|
(405) |
where and are independent of
but can depend on , then Eq. (404) implies
|
|
|
(406) |
Using Eq. (405), we conclude
|
|
|
(407) |
which can be a nonlinear equation.
This means that if is negligible, then and
obey the same equation. In our model, the equation of
motion for (see Eq. (279)) makes
|
|
|
(408) |
in Eq. (407) which upon expansion gives
|
|
|
(409) |
matching Eq. (93). The in this system
is
|
|
|
(410) |
Because of the mismatch of between and ,
we cannot conclude that is constant from
this argument alone.
We will now show that the conservation equation from symmetry
together with a mild assumption about the lack of resonance allows
one to conclude that is approximately
frozen during and after the transition. Start with the linear perturbation
equation for the the current conservation
|
|
|
(411) |
where we have defined
|
|
|
(412) |
and gone to Fourier space. Let be the first time in the
time-independent conformal era when the term can be neglected
in Eq. (411). We can conclude that
which is set during the time-independent conformal era
to be
|
|
|
|
(413) |
is conserved while our quantity of interest
is related to this constant through
|
|
|
(414) |
coming from Eq. (412). Integrating,
we find
|
|
|
|
(415) |
|
|
|
|
(416) |
where is the time at which
starts to change in time (i.e. deviate from the conformal behavior).
Because the lighter energy mode becomes purely
the after the settles to the minimum of
the potential, we know
asymptotically. Hence, for we shall assume
|
|
|
(417) |
Additionally we consider a smooth non-resonant adiabatic transition
of the background radial field such that
|
|
|
(418) |
because the equation of motion near the transition time can be solved
to obtain
|
|
|
(419) |
which shows that the coefficient of is suppressed.
Hence, compared to the first term, the second term in the integral
falls off rapidly for . Thus, we simplify the integral
as
|
|
|
|
(420) |
Using the conservation equation, this becomes
|
|
|
|
(421) |
where we used Eq. (418). Since we can solve during
the time-independent conformal era
|
|
|
(422) |
we obtain the relation
|
|
|
|
(423) |
This indicates that in the long wavelength limit, the isocurvature
perturbation is conserved for modes outside of the horizon at the
transition time even in the presence of a large rotating background.