A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y...

14
A ΛCDM Bounce Scenario Yi-Fu Cai 1, * and Edward Wilson-Ewing 2, 3, 1 Department of Physics, McGill University, Montr´ eal, QC H3A 2T8, Canada 2 Department of Physics and Astronomy, Louisiana State University, Baton Rouge, 70803 USA 3 Max Planck Institute for Gravitational Physics (Albert Einstein Institute), Am M¨ uhlenberg 1, 14476 Golm, Germany, EU We study a contracting universe composed of cold dark matter and radiation, and with a positive cosmological constant. As is well known from standard cosmological perturbation theory, under the assumption of initial quantum vacuum fluctuations the Fourier modes of the comoving curvature perturbation that exit the (sound) Hubble radius in such a contracting universe at a time of matter- domination will be nearly scale-invariant. Furthermore, the modes that exit the (sound) Hubble radius when the effective equation of state is slightly negative due to the cosmological constant will have a slight red tilt, in agreement with observations. We assume that loop quantum cosmology captures the correct high-curvature dynamics of the space-time, and this ensures that the big-bang singularity is resolved and is replaced by a bounce. We calculate the evolution of the perturba- tions through the bounce and find that they remain nearly scale-invariant. We also show that the amplitude of the scalar perturbations in this cosmology depends on a combination of the sound speed of cold dark matter, the Hubble rate in the contracting branch at the time of equality of the energy densities of cold dark matter and radiation, and the curvature scale that the loop quantum cosmology bounce occurs at. Importantly, as this scenario predicts a positive running of the scalar index, observations can potentially differentiate between it and inflationary models. Finally, for a small sound speed of cold dark matter, this scenario predicts a small tensor-to-scalar ratio. PACS numbers: 98.80.Qc, 98.80.Cq I. INTRODUCTION Observations of the cosmic microwave background (CMB) —most recently [1, 2]— have clearly established that scalar perturbations in the early universe were nearly scale-invariant. It is thus necessary for any re- alistic cosmological model to generate, in some fashion, scale-invariant perturbations. To achieve this, many cosmological models rely on the presence of matter fields (typically scalar fields) that have not yet been observed in nature. The new matter fields are necessary in these models as they play an essential role in the generation of scale-invariant perturbations. While it is of course a requirement for any cosmological scenario to predict near scale-invariance in order to be potentially viable, there are some cosmological scenarios where it is possible to avoid the weakness of postulating the existence of unknown matter fields and nonetheless obtain scale-invariance. We shall study one such model in this paper. This cos- mological model consists of a spatially flat Friedmann- Lemaˆ ıtre-Robertson-Walker (FLRW) universe, with a positive cosmological constant, cold dark matter (CDM), and radiation. These are three ingredients known to be present in our universe, and we will not assume the exis- tence of any other matter fields. We also assume that the initial conditions are such that the space-time curvature is small and the universe is large and contracting. As the universe contracts, the space-time curvature * [email protected] [email protected] will increase, and quantum gravity effects are expected to become important at some point, likely when the space- time curvature nears the Planck scale. In this work, we will assume that loop quantum cosmology (LQC) cap- tures the salient non-perturbative quantum gravity ef- fects in the very early universe. LQC, a mini-superspace approach to quantum cosmology motivated by loop quan- tum gravity, predicts that a bounce occurs near the Planck scale and that, once these quantum gravity ef- fects are included, the space-time is free of the singular- ities that appear in classical general relativity [3–5]. Thus, this model will be that of a bouncing universe, with a matter content of radiation and cold dark mat- ter and a positive cosmological constant. Now, it is well known in cosmological perturbation theory that, for perturbations that are initially in the quantum vacuum state, the Fourier modes that reach the long wavelength limit in a contracting space-time whose dynamics are dominated by a pressureless matter field become scale- invariant [6, 7]. (Note that the long-wavelength limit of a Fourier mode does not always coincide with the mode exiting the Hubble radius — here the relevant length scale is the sound Hubble radius, as explained in more detail later.) Furthermore, if the pressure is slightly neg- ative, for example due to the presence of a positive cos- mological constant, then the long wavelength perturba- tion modes will be almost scale-invariant with a slight red tilt. Therefore, in the model considered here, we ex- pect the modes that become large during the epoch of the universe that is dominated by cold dark matter to be almost scale-invariant, and those that become large when the effective equation of state is slightly negative to have a small red tilt. arXiv:1412.2914v2 [gr-qc] 28 Jan 2015

Transcript of A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y...

Page 1: A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y 1Department of Physics, McGill University, Montr eal, QC H3A 2T8, Canada 2Department of Physics

A ΛCDM Bounce Scenario

Yi-Fu Cai1, ∗ and Edward Wilson-Ewing2, 3, †

1Department of Physics, McGill University, Montreal, QC H3A 2T8, Canada2Department of Physics and Astronomy, Louisiana State University, Baton Rouge, 70803 USA

3Max Planck Institute for Gravitational Physics (Albert Einstein Institute), Am Muhlenberg 1, 14476 Golm, Germany, EU

We study a contracting universe composed of cold dark matter and radiation, and with a positivecosmological constant. As is well known from standard cosmological perturbation theory, under theassumption of initial quantum vacuum fluctuations the Fourier modes of the comoving curvatureperturbation that exit the (sound) Hubble radius in such a contracting universe at a time of matter-domination will be nearly scale-invariant. Furthermore, the modes that exit the (sound) Hubbleradius when the effective equation of state is slightly negative due to the cosmological constant willhave a slight red tilt, in agreement with observations. We assume that loop quantum cosmologycaptures the correct high-curvature dynamics of the space-time, and this ensures that the big-bangsingularity is resolved and is replaced by a bounce. We calculate the evolution of the perturba-tions through the bounce and find that they remain nearly scale-invariant. We also show that theamplitude of the scalar perturbations in this cosmology depends on a combination of the soundspeed of cold dark matter, the Hubble rate in the contracting branch at the time of equality of theenergy densities of cold dark matter and radiation, and the curvature scale that the loop quantumcosmology bounce occurs at. Importantly, as this scenario predicts a positive running of the scalarindex, observations can potentially differentiate between it and inflationary models. Finally, for asmall sound speed of cold dark matter, this scenario predicts a small tensor-to-scalar ratio.

PACS numbers: 98.80.Qc, 98.80.Cq

I. INTRODUCTION

Observations of the cosmic microwave background(CMB) —most recently [1, 2]— have clearly establishedthat scalar perturbations in the early universe werenearly scale-invariant. It is thus necessary for any re-alistic cosmological model to generate, in some fashion,scale-invariant perturbations.

To achieve this, many cosmological models rely on thepresence of matter fields (typically scalar fields) that havenot yet been observed in nature. The new matter fieldsare necessary in these models as they play an essentialrole in the generation of scale-invariant perturbations.While it is of course a requirement for any cosmologicalscenario to predict near scale-invariance in order to bepotentially viable, there are some cosmological scenarioswhere it is possible to avoid the weakness of postulatingthe existence of unknown matter fields and nonethelessobtain scale-invariance.

We shall study one such model in this paper. This cos-mological model consists of a spatially flat Friedmann-Lemaıtre-Robertson-Walker (FLRW) universe, with apositive cosmological constant, cold dark matter (CDM),and radiation. These are three ingredients known to bepresent in our universe, and we will not assume the exis-tence of any other matter fields. We also assume that theinitial conditions are such that the space-time curvatureis small and the universe is large and contracting.

As the universe contracts, the space-time curvature

[email protected][email protected]

will increase, and quantum gravity effects are expected tobecome important at some point, likely when the space-time curvature nears the Planck scale. In this work, wewill assume that loop quantum cosmology (LQC) cap-tures the salient non-perturbative quantum gravity ef-fects in the very early universe. LQC, a mini-superspaceapproach to quantum cosmology motivated by loop quan-tum gravity, predicts that a bounce occurs near thePlanck scale and that, once these quantum gravity ef-fects are included, the space-time is free of the singular-ities that appear in classical general relativity [3–5].

Thus, this model will be that of a bouncing universe,with a matter content of radiation and cold dark mat-ter and a positive cosmological constant. Now, it iswell known in cosmological perturbation theory that, forperturbations that are initially in the quantum vacuumstate, the Fourier modes that reach the long wavelengthlimit in a contracting space-time whose dynamics aredominated by a pressureless matter field become scale-invariant [6, 7]. (Note that the long-wavelength limit ofa Fourier mode does not always coincide with the modeexiting the Hubble radius — here the relevant lengthscale is the sound Hubble radius, as explained in moredetail later.) Furthermore, if the pressure is slightly neg-ative, for example due to the presence of a positive cos-mological constant, then the long wavelength perturba-tion modes will be almost scale-invariant with a slightred tilt. Therefore, in the model considered here, we ex-pect the modes that become large during the epoch ofthe universe that is dominated by cold dark matter tobe almost scale-invariant, and those that become largewhen the effective equation of state is slightly negativeto have a small red tilt.

arX

iv:1

412.

2914

v2 [

gr-q

c] 2

8 Ja

n 20

15

Page 2: A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y 1Department of Physics, McGill University, Montr eal, QC H3A 2T8, Canada 2Department of Physics

Various realizations of the matter bounce scenario sug-gested in [6, 7] have been considered in the literature, butin most studies the matter content is taken to be scalarfields with specific potentials that can mimic pressurelessmatter fields for some specific initial conditions. To thebest of our knowledge, this is the first study of a mat-ter bounce scenario where the pressureless matter fieldis taken to be cold dark matter and that the effect of apositive cosmological constant is also included. We willcompare the predictions of this scenario with other real-izations of the matter bounce scenario in Sec. V.

In this paper we calculate the spectrum of the cosmo-logical perturbations for this ΛCDM bounce scenario. Bythe use of some approximations, it is possible to completethe calculations entirely analytically; we also solve theequations numerically in order to provide a check on thevalidity of the approximations. In Sec. II we study thedynamics of the background, first analytically and thennumerically. Then in Sec. III we calculate the evolutionof the scalar perturbations, from their initial quantumvacuum state to their final form after the bounce in thebackground FLRW space-time. Once again, this calcula-tion is first done analytically with the help of some ap-proximations, and is solved numerically afterwards. Wecontinue in Sec. IV by determining the spectrum of theprimordial gravitational waves, and end in Sec. V with adiscussion. We use units where c = 1, but keep G and~ explicit except where stated otherwise, typically in thesections devoted to the numerical studies.

II. HOMOGENEOUS BACKGROUND

As explained in the Introduction, we are interested instudying the dynamics of a flat FLRW cosmology with apositive cosmological constant Λ and whose matter con-tent is composed of radiation and cold dark matter, whichare modeled as perfect fluids.

Classically, the dynamics are given by the Friedmannequations, while in LQC there exists a Hamiltonianconstraint operator that generates the evolution of thewave function representing the quantum cosmology state.While the full quantum evolution is in general rathercomplicated, for sharply peaked states (i.e., states thatadmit a clear semi-classical interpretation at low curva-ture scales) the full quantum dynamics are very well ap-proximated by a set of effective equations [8–10]. Thekey point is that since the state is sharply peaked (and itremains sharply peaked throughout the entire evolution,including at the bounce point), it is meaningful to speakof an effective geometry, with an effective scale factor,and to ask what equations of motion govern the dynam-ics of this effective scale factor; these equations are calledthe effective equations. As radiation will dominate thedynamics in the high-curvature regime, it is enough toconsider the effective equations for a radiation-dominated

flat FLRW space-time, which are given by

H2 =8πG

(1− ρ

ρc

), (1)

where H = a/a is the Hubble rate (in proper time), ρ isthe energy density of the radiation matter field, and ρc ∼ρPl is the critical energy density, which is of the order ofthe Planck energy density. In addition, the matter fieldsatisfied the continuity equation

ρ+ 4Hρ = 0, (2)

where we have used the fact that the pressure of a ra-diation perfect fluid is P = ρ/3. Note that the classicalFriedmann equations are obtained in the limit ρc → ∞.In this paper, we will restrict our analysis to the effec-tive equations of LQC, but a full quantum treatment of aradiation-dominated space-time in LQC is given in [11].

In the first part of this section, we will use some rea-sonable approximations in order to derive some analyt-ical results, and in the second section we present somenumerical results that in part complement the analyticresults and in part provide a check on the validity of theapproximations.

To be specific, we shall make three approximationsin the analytical treatment of the background: first, weshall assume that quantum gravity effects are negligiblealready a few orders of magnitude away from the LQCbounce. This has been verified in numerical simulations[11], and can also be seen from studying the effectiveequations. This will allow us to solve the classical Fried-mann equations away from the bounce, and the LQC cor-rections will only become relevant when the space-timecurvature nears the Planck scale during the radiation-dominated epoch.

Second, we will assume that the evolution of the back-ground universe can be broken into two distinct eras: thefirst one which is dominated by the combination of thecosmological constant and cold dark matter, and anotherera which is dominated by radiation. We assume a dis-continuous change in the equation of state between thesetwo eras, and impose continuity in the scale factor andin the (conformal) Hubble rate during this transition.This approximation is supported by results in Sec. II Bthat show that the transition between the matter- andradiation-dominated epochs occurs very rapidly.

Finally, recall that the goal of this paper is to calcu-late the power spectrum of the perturbations. As is wellknown, especially from calculations in inflation, the keyingredient that determines the scale-dependence of theperturbations is the equation of state of the backgroundat the time that the mode reaches the long wavelengthlimit. Therefore, in order to simplify calculations, wewill assume a constant equation of state during the time-frame that the perturbation modes of interest reach thelong wavelength limit (i.e., when the dynamics of thespace-time are dominated by the CDM, but the cosmo-logical constant provides a small correction to the effec-tive equation of state). This approximation is justified by

2

Page 3: A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y 1Department of Physics, McGill University, Montr eal, QC H3A 2T8, Canada 2Department of Physics

the effective equation of state being nearly constant dur-ing the period of interest, as seen in Sec. II B. Nonethe-less, one should keep in mind that the effective equationof state —due to the combination of Λ and CDM— isin fact changing in time, and in general will be slightlydifferent for different modes. As we shall see later, thiseffect leads to a running of the scalar index ns.

A. Analytic Treatment

Using the effective equations, we shall first determinethe dynamics around the bounce point, and then solvefor the scale factor at earlier pre-bounce times when thespace-time curvature is much smaller.

1. Radiation-Dominated Epoch

It is easy to see that (2) implies that

ρ(t) =ρoa(t)4

, (3)

where ρo is a constant of integration, and this can beused to solve (1), giving

a(t) =

(32πGρo

3(t− to)2 +

ρoρc

)1/4

,

where to is another constant of integration. It is of coursepossible to choose any values for to and ρo; for conve-nience we shall set to = 0 so that the bounce occurs att = 0, and ρo = ρc so that the value of the scale factorat the bounce point is 1. Then,

a(t) =

(32πGρc

3t2 + 1

)1/4

. (4)

Well before and after the bounce (|t| √

3/32πGρc),the space-time curvature is much smaller than the Planckscale and the scale factor is very well-approximated bythe classical solution

a(t) = a1/4o

√|t|, (5)

where we have defined

ao =32πGρc

3(6)

for later convenience, and the Hubble rate is given by

H =1

2t. (7)

It is also easy in the classical regime to change to con-formal time η via the relation adη = dt, which gives

|t| =√

2πGρc3

η2, (8)

which in turn shows that the scale factor in terms ofconformal time is given by

a(η) =

√8πGρc

3|η|, (9)

and the conformal Hubble rate, again in the classicalregime, is given by

H =1

η= −

(8πGρc

3

)1/4√−H; (10)

the second equality holds for the contracting epoch of thecosmology where H < 0 and H < 0.

2. Cold Dark Matter and Λ

Now we shall consider the earlier epoch where colddark matter and the cosmological constant dominate thedynamics (the ΛCDM era). In this regime, the classicalFriedmann equations in conformal time can be writtenas

H2 =8πG

3a2 (ρCDM + ρΛ) =

8πG

3a2 ρtot, (11)

where ρΛ = Λ/8πG, and the combined continuity equa-tion (also in conformal time) for cold dark matter andthe cosmological constant is given by

ρ′tot + 3H (ρtot + Ptot) = 0. (12)

Since PΛ = −ρΛ and PCDM = 0, it follows that theeffective equation of state for cold dark matter and thecosmological constant combined1 is Ptot = ωρtot, with−1 ≤ ω ≤ 0.

In order to solve these two equations exactly, we shallassume that ω = −δ is a constant, and furthermore, sincewe are interested in the regime where the dynamics aredominated by the cold dark matter, we also take δ 1.Of course, the effective equation of state does not re-main constant in this setting, but recall that we are inter-ested in calculating the power spectrum of cosmologicalperturbations, and their scale-dependence depends mostsensitively on the effective equation of state at the timewhen they reach the long wavelength limit. Therefore,the calculations where the spectra of the scalar and ten-sor perturbations are determined are to be understood asbeing for the modes that reach the long wavelength limitwhen the effective equation of state is given by ω = −δ,and so the specific value of δ will vary from one mode

1 In fact, we expect that PCDM = ε2ρCDM with 0 < ε 1, buthere we are interested in the situation where the small positivecontribution to ω from the cold dark matter and the small neg-ative contribution to ω from the cosmological constant combineto give a slightly negative ω.

3

Page 4: A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y 1Department of Physics, McGill University, Montr eal, QC H3A 2T8, Canada 2Department of Physics

to another. Note that this variation will be monotonicwith δ becoming closer and closer to zero for shorter andshorter wavelengths, or for larger and larger k. The ex-act rate at which this occurs will depend on the relativecontributions of cold dark matter and the cosmologicalconstant to the total matter energy density.

In the approximation that δ is constant, the total en-ergy density behaves as

ρtot =ρeffa3(1−δ) , (13)

where ρeff is a constant of integration, and the scalefactor is given by

a(η) =

[√2πGρeff

3(1− 3δ)(η − ηo)

]2/(1−3δ)

, (14)

where ηo is also a constant of integration. It follows thatthe conformal Hubble rate is

H =2

(1− 3δ)(η − ηo). (15)

In order to determine the values of ρeff and ηo, letus assume that the transition between the radiation-dominated epoch and the ΛCDM era occurs at the equal-ity conformal time ηe. Then, imposing that the scalefactor and the conformal Hubble rate be continuousat the transition time, we find that ρeff = ρc/a

1+3δe

and ηo = ηe − 2/[(1 − 3δ)He], where ae = a(ηe) andHe = H(ηe) respectively. Then, the scale factor can berewritten as

a(η) = ae

(η − ηoηe − ηo

)2/(1−3δ)

. (16)

3. Summary

Thus, for early times η ≤ ηe, the scale factor is givenby (16); then for ηe ≤ η the relation (9) holds so long asquantum-gravity effects are negligible. When quantumgravity effects become important, it is necessary to use(4) to describe the dynamics of the scale factor.

B. Numerics of the Background Dynamics

In this subsection we numerically study the back-ground solution presented in the ΛCDM model within thecontext of LQC. To be explicit, we study the backgrounduniverse by separating the evolution into two periods.In the first stage, we apply the realistic data from thePlanck results and numerically solve for the period thatthe contracting universe evolves from dark energy domi-nation to the matter-dominated phase. In particular, weassume that the density parameters of all matter compo-nents are the same as what we observed today, which are

given by Ωm ≡ ρm/ρtot = 0.314, ΩΛ ≡ ρΛ/ρtot = 0.686and a deduced value Ωr ≡ ρr/ρtot = 9.23× 10−5, respec-tively [2]. During this phase, the numerical result of thebackground evolution is provided in Fig. 1.

One particularly important result is shown in the insetof Fig. 1a: the effective equation of state evolves veryslowly as the era of matter-domination is approached.This shows that, for modes that reach the long wave-length limit in this time-frame, any effect due to an evolv-ing effective equation of state will be very small, and itis justified to make the approximation that δ is constantin the analytic calculations.

Then, in the second stage, we consider a toy modelof a universe filled with a dark matter component and aradiation component with their energy densities evolvingas follows,

ρm = ρim

(aia

)3(1−δ), ρr = ρir

(aia

)4

, (17)

respectively, where ρim and ρir are their initial valueswhen a = ai. Note that while it would be nice to si-multaneously include both matter fields and a cosmolog-ical constant, this is significantly more expensive froma computational point of view. These numerical studiespresented here already provide strong support for the ap-proximations used in the analytic section, and we leavemore detailed numerical studies for future work.

In order to relate this stage of the dynamics with theprevious one, the pressure is taken to have a very small(but non-vanishing) negative value. Also, to make thecomparison between the analytic and numerical resultsas simple as possible, we adopt the conformal time η andset the value of the scale factor a to be unity at thebouncing moment tB = 0 in the numerical calculation.The dynamics are calculated by numerically solving thesecond Friedmann equation H′ instead of the first one in(11), which is given by

H′ = −4πG

3a2

(ρtot + 3Ptot −

4ρ2tot

ρc− 6ρtotPtot

ρc

),(18)

in the case of LQC. In addition, the continuity equa-tion (12) is applied so that the background equationsof motion are self-complete after an initial value of thebackground energy density has been imposed.

For the numerics, we work in units of the reducedPlanck mass MPl ≡ 1/

√8πG (with ~ = 1) for all model

parameters with dimensions. As an explicit example (al-though not a realistic one), we choose the values of theenergy densities, the critical density ρc and an effectiveequation of state parameter for the CDM at the initialmoment to be

ρim = 1.1× 10−24 , ρir = 5.1× 10−28 ,

ρc = 2.9× 10−9 , δ = 0.05 , (19)

and our numerical results are shown in Fig. 2.The evolution of the scale factor, which is depicted

by a blue solid curve in the left panel, explicitly shows

4

Page 5: A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y 1Department of Physics, McGill University, Montr eal, QC H3A 2T8, Canada 2Department of Physics

- 1 4 - 1 2 - 1 0 - 8 - 6 - 4 - 2 0

- 0 . 8

- 0 . 4

0 . 0

0 . 4

0 . 8

- 2 . 0 - 1 . 6 - 1 . 2 - 0 . 8 - 0 . 4- 0 . 0 8- 0 . 0 40 . 0 00 . 0 40 . 0 8

t ( G y r )

w = p / ρ

(a)

- 1 4 - 1 2 - 1 0 - 8 - 6 - 4 - 2 00 . 0

0 . 2

0 . 4

0 . 6

0 . 8

1 . 0

t ( G y r )

Ω m ΩΛ

(b)

FIG. 1. Evolutions of the background equation of state parameter ω (the purple solid curve in the left panel), and the density parametersΩi ≡ ρi/ρtot (the red dotted and the blue dashed lines in the right panel) as a function of the cosmic time t (in units of per billionyears) in the model under consideration. The horizontal axis denotes the cosmic time t. The initial values of background parametersare assumed to be the same as today’s universe.

- 2 x 1 0 1 0 - 1 x 1 0 1 0 0 1 x 1 0 1 0 2 x 1 0 1 00

4

8

1 2

1 6

η

a

(a)

- 2 x 1 0 1 0 - 1 x 1 0 1 0 0 1 x 1 0 1 0 2 x 1 0 1 0

- 2 . 0 x 1 0 - 5

- 1 . 0 x 1 0 - 5

0 . 0

1 . 0 x 1 0 - 5

2 . 0 x 1 0 - 5

η

(b)

- 1 x 1 0 1 2 0 1 x 1 0 1 2

0 . 0

0 . 2

0 . 4

0 . 6

0 . 8

1 . 0

η

Ω m Ω r

(c)

FIG. 2. Evolutions of the scale factor a (the blue solid curve in the left panel), the conformal Hubble parameter H (the red dotted curve inthe middle panel) and the density parameters Ωi ≡ ρi/ρtot (the green dash-dotted and the orange dashed lines in the right panel) as afunction of the conformal time η in the model under consideration. The horizontal axis denotes the conformal time η. The backgroundparameters chosen for the numerics are given in Eq. (19).

that a non-singular bouncing solution is obtained in ourmodel due to the quantum gravity effects captured byLQC, in particular, the minimal value of the scale factoris non-zero. From the middle panel, one can read moredetails about the background evolution. For example, theabsolute value ofH is increasing when η is about less than−2× 109. After that, H becomes approximately a linearfunction of the conformal time and correspondingly, theuniverse enters the bouncing phase with H evolving fromthe negative valued regime to the positive valued one.Eventually, the value of H decreases after η ≈ 2 × 109

and hence the universe naturally connects to a regularthermal expanding phase after the bounce; there is noneed for reheating.

The right panel of Fig. 2 characterizes the evolutions ofthe density parameters of the dark matter and radiationin our model, which are defined by

Ωm(r) ≡ρm(r)

ρtot, (20)

with the subscripts “m” and “r” representing dark mat-ter and radiation respectively. One can see that the uni-verse was originally dominated by the cold dark matterwith Ωm ' 1, as described the green dash-dotted curve.During the cosmic contraction the contribution of radia-tion, which is depicted by the orange dashed line, growsfaster than that of dark matter and then dominates overthe background evolution before the bounce. After thebounce, the universe would have experienced a period ofradiation-dominated expanding phase and eventually en-ters the CDM era and hence is in qualitative agreementwith cosmological observations. Note that the sharptransition between these two eras provides justificationfor the assumption of a discontinuous transition betweenmatter- and radiation-domination used in the analyticcalculations. Also, it is important to keep in mind that amore careful choice of the initial parameters would makethe model more precisely consistent with experimentaldata.

5

Page 6: A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y 1Department of Physics, McGill University, Montr eal, QC H3A 2T8, Canada 2Department of Physics

III. SCALAR PERTURBATIONS

In this section, we will calculate the final spectrumof scalar perturbations after the bounce, assuming theybegin in the quantum vacuum state in the distant pastof the pre-bounce epoch.

In cosmological perturbation theory, it is convenient touse the gauge-invariant Mukhanov-Sasaki variable [12]

v = zR, (21)

where R is the comoving curvature perturbation and

z =a√ρ+ P

csH. (22)

Linear perturbations can be handled in LQC by followingthe ‘separate universe’ approach presented in [13, 14],and from the resulting quantum theory it is possible toderive effective equations that can be used to calculateexpectation values of sharply-peaked states. The LQCeffective equations for the Mukhanov-Sasaki variable are[15, 16]

v′′ − c2s(

1− 2ρ

ρc

)∇2v − z′′

zv = 0, (23)

and it is easy to see that the standard classical expres-sion is recovered in the limit ρc → ∞. The effectiveequation (23) is expected to provide a good approxima-tion to the full quantum dynamics for modes that alwaysremain large compared to the Planck length [9], whichis the case for the observationally relevant modes in thematter bounce scenario.

In this section, using (23) we will determine how theFourier modes vk evolve, from their initial quantum vac-uum state, to their form as they exit the sound Hub-ble radius2 rsH = cs/H and finally how they propagatethrough the bounce. As we shall see, the Fourier modesthat exit the sound Hubble radius during the periodof matter domination in the contracting branch becomescale-invariant and therefore these modes are of particu-lar interest. This is why in this paper we will only con-sider this family of the Fourier modes vk, and ignore themodes that exit the Hubble radius either before (duringthe epoch dominated by the cosmological constant) orafter (during the radiation-dominated phase).

In the first part, we present analytical calculations, andin the second, numerical simulations. As in the previoussection, it is necessary to make certain approximations inorder to make the analytical calculations tractable andthe numerical studies in the second part serve in part tocheck that the approximations are valid.

2 Since the sound speed of the matter fields in this model is not 1,we find that the Fourier modes reach the long wavelength limitwhen the mode exits the sound Hubble radius, not the Hubbleradius. These quantities differ by a factor of cs. Note that it isequivalent to state that a given mode reaches the long wavelengthlimit when its ‘sound wavelength’ exits the Hubble radius.

A. Analytic Treatment

We will begin by solving the dynamics of the pertur-bations in the ΛCDM era —and the modes of interestare those that reach the long wavelength limit duringthis era— and then determine their evolution duringthe radiation-dominated epoch, including through thebounce.

In order to solve the equations of motion for vk, itis necessary to make the following approximations: (i)we only consider the modes that reach the long wave-length limit during the ΛCDM era, where the back-ground matter field is modelled as a perfect fluid witha small and negative constant equation of state as ex-plained in Sec. II A 2; (ii) we assume continuity in vk andv′k at the spatial slice where we approximate the equa-tion of state changing discontinuously to the radiation-dominated epoch; and (iii) we work in the long wave-length limit during the bounce period.

Recall from Sec. II B that the equation of state changesvery slowly in the regime where the effective equation ofstate is slightly negative. This supports the first approx-imation, since the key ingredient in determining the longwavelength spectrum of scalar perturbations is the equa-tion of state of the background at the time that the givenmode exits the sound Hubble radius. Since the effectiveequation of state is changing slowly, we expect correctionsto the approximation of a constant equation of state tobe subleading. Finally, the validity of approximations (ii)and (iii) is verified numerically in Sec. III B.

1. The ΛCDM Era

For the contracting portion of the space-time wherethe scale factor is given by (16) (and safely neglectingquantum gravity effects at this stage), (23) is

v′′k + c2s k2vk −

2(1 + 3δ)

(1− 3δ)2(η − ηo)2vk = 0. (24)

Since the sound speed of cold dark matter is unknown,we set cs = ε which we assume to be constant. We expectε to be a small positive number.

The solutions to this differential equation are

vk =√−(η − ηo)

(A1H

(1)n [−εk(η − ηo)]

+A2H(2)n [−εk(η − ηo)]

), (25)

where H(1)n and H

(2)n are the Hankel functions, and

n =

√2(1 + 3δ)

(1− 3δ)2+

1

4≈ 3

2+ 6δ +O(δ2), (26)

where after the last equality we drop terms of order δ2

and higher (recall that δ 1).

6

Page 7: A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y 1Department of Physics, McGill University, Montr eal, QC H3A 2T8, Canada 2Department of Physics

Choosing the initial conditions to be quantum vacuumfluctuations sets A1 =

√π~/4 and A2 = 0.

Then, as η approaches ηe, some modes satisfy −εk(η−ηo) 1. These modes are said to be in the long wave-length limit, and in this limit it is possible to use the smallargument expansion of the Hankel functions to show that

vk =

√−π~(η − ηo)

4

[(εk)n

Γ(n+ 1)

(−(η − ηo)

2

)n− i Γ(n)

π(εk)n

(−2

η − ηo

)n ](27)

=

√8~9

(εk)3/2+6δH−(2+6δ)

− i√

~4

(εk)−3/2−6δH1+6δ, (28)

where in the second equality the time dependence hasbeen rewritten in terms of the conformal Hubble rate,and the exponents are accurate to first order in δ, whilethe numerical prefactors are only accurate to zeroth or-der in δ. It is straightforward to determine higher ordercorrections in δ, but this will not be necessary here.

2. Radiation-Dominated Epoch

During the radiation-dominated epoch, the scale factoris proportional to η while the sound speed is given by1/√

3, and so the Mukhanov-Sasaki variable satisfies theequation

v′′k +k2

3vk = 0, (29)

at least in the classical regime where quantum gravityeffects are negligible. Note that due to the drastic changein the speed of sound, some modes that were in the longwavelength regime may at first be in the short wavelengthlimit at the onset of the radiation-dominated epoch.

Therefore, the relevant solutions are

vk = B1 sinkη√

3+B2 cos

kη√3, (30)

and B1 and B2 can be determined from (27) by demand-ing that vk and v′k be continuous at ηe. Note that asthe bounce is approached η → 0 and therefore the sec-ond term with the prefactor B2 will dominate, so we candrop B1. Imposing continuity in vk and v′k gives

B2 =− i√

~4

(εk)−3/2−6δ coskηe√

3H1+6δe

− i√

3~16

(εk)−3/2−6δk−1 sinkηe√

3H2+6δe . (31)

Some modes will have rebecome short wavelength modesdue to the drastic change in the sound speed. For these

modes, the small argument expansion for the trigonemet-ric functions cannot be used, and these terms will notbe scale-invariant. Thus, we expect that scale-invariancewill only be obtained for the modes that satisfy kηe 1.We will return to this point later.

3. The Bounce

In the contracting phase, as the bounce is approached(but before quantum gravity effects become important)we have |kη| 1 and in this limit the solution for theMukhanov-Sasaki variable (30) tends to

vk = B2, (32)

with B2 given in (31).During the bounce, all of the modes of cosmological in-

terest remain in the long-wavelength limit, and thereforethe equation of motion for vk is

v′′k −z′′

zvk = 0, (33)

where z = a√ρ+ P/csH = 4

√ρc a

3/(ao t) [recall that

cs = 1/√

3 during radiation-domination and ao is definedin (6)], and the solution is

vk = C1z + C2z

∫η

z(η)2. (34)

Note that z is not simply proportional to a due to thequantum gravity effects that modify the Friedmann equa-tion at high curvatures as seen in (1). The integral canbe evaluated by rewriting it in terms of the proper timevia the relation dt = adη, giving

vk = C1 z + C2 z t

[2F1

(12 ,

34 ; 3

2 ,−aot2)− 1

a3

], (35)

where C2 has been redefined in order to absorb somenumerical factors.

In the classical pre-bounce era (t −√ao), this ex-pression must agree with (32) and this uniquely deter-mines

C1 = π

√G

6

Γ(

14

)Γ(

34

) B2, C2 = −√

2πGao3

B2. (36)

Note that during this calculation it is important to keepin mind that t = −|t| in the pre-bounce era.

It is also easy to calculate the form of the scalar per-turbations in the classical post-bounce era by taking thelimit t √ao in (35), which gives in terms of the co-moving curvature perturbation

Rk =vkz

= 2C1 +O(t−1), (37)

where we have only kept the dominant contribution,namely the constant mode which is the only one thatdoes not decay with time.

7

Page 8: A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y 1Department of Physics, McGill University, Montr eal, QC H3A 2T8, Canada 2Department of Physics

4. Results

The amplitude of the comoving curvature perturbationof 2C1 after the bounce depends on B2 via (36) and soit is easy to check whether the resulting scalar perturba-tions (37) are scale-invariant or not. Since the only de-pendence of k in C1 resides in B2, a quick examination of(31) suffices to determine the scale-dependence of R. Asone can readily verify, one obtains near scale-invarianceonly in the limit of |kηe| 1 (otherwise the dominantcontribution would be oscillations superimposed over ared spectrum), in which case

B2 = −3 i

4

√~ (εk)−

32−6δH1+6δ

e

(1 +O

(k2η2

e

)). (38)

Therefore, a necessary condition for scale-invariance isthat

|kηe| 1, (39)

that is to say that the modes that become (nearly) scale-invariant during the matter-dominated contracting eramust remain outside the sound Hubble radius duringthe entire contracting radiation-dominated epoch and thebounce in order to remain (nearly) scale-invariant.

For these modes, the power spectrum is

∆2R =

k3

2π2|R|2

=

√3π

2

(Γ(

14

)Γ(

34

))2 √ρcρPl· |He|`Pl

ε3

×(

8πGρc|He|3k4

)3δ

, (40)

where ρPl = 1/(G2~) and `Pl =√G~, and the tilt is

given by

ns = 1− 12 δ. (41)

Thus, the observation of the tilt to be ns ≈ 0.96 [1, 2] setsδ ≈ 0.003, which means that when the wavelength k−1

exits the sound Hubble radius ε/H the effective equationof state must have been ωeff ≈ −0.003.

Also, in this model we predict a small running of thescalar index. This is due to the following two effects: (i)the departure from scale-invariance in a small interval ofk depends on the background effective equation of state,and (ii) the background equation of state is dynamical.The smallest values of k reach the long-wavelength limitfirst, at a time when Λ contributes slightly more to thebackground dynamics than it does at later times whenlarger values of k reach the long-wavelength limit. There-fore, as k increases, the background equation of state atthe ‘sound-Hubble-crossing’ time also increases,

dωeff

dk> 0, (42)

and since ωeff = −δ and ns = 1− 12δ, it follows that

dnsdk

> 0. (43)

Therefore another prediction of this realization of thematter bounce scenario is for the scalar index ns to in-crease with k. Although the presence of this effect isclear in this model, its amplitude is not known. In orderto calculate the expected amplitude of dns/dk, it wouldbe necessary to know quite precisely the energy densitiescorresponding to cold dark matter and the cosmologicalconstant during the contracting phase, which is not aneasy task especially since we do not expect the universeto be symmetric around the bounce point, as we shalldiscuss next.

Nonetheless, the qualitative result (43) is a clear pre-diction for this realization of the matter bounce scenario.

Finally, as stated above, the results (40) and (41) onlyhold for Fourier modes that satisfy the condition (39).In order to better understand this condition, it is usefulto rewrite it in terms of the physical wave number k =a(t) · kphys(t) and of the Hubble rate at equality He via(5), (7) and (10),

kphys(te)

|He| 1. (44)

A good choice to ensure that kphys corresponds to modesthat are observed in the cosmic microwave backgroundtoday, is to choose3 k? = 0.05 Mpc−1 = 10−59`−1

Pl , whichlies roughly in the middle of the logarithmic range ofthe scales probed by the Planck telescope [2]. The valueof k? at the time of equality is given by the relationk?(te) = a(to) · k?(to)/a(te). With our choice of con-ventions of setting a(t = 0) = 1 at the bounce, it followsthat a(to) ∼ 1031. This can be calculated from the factthat (i) the scale factor increased by a factor of ∼ 104 af-ter matter-radiation equality until today, and (ii) that inthe expanding branch matter-radiation equality is knownto occur at t+e ∼ 104 years ∼ 6 × 1054tPl giving a scale

factor of a(t+e ) = a1/4o√te ∼ 1027 (assuming ρc ∼ ρPl)

at the time of matter-radiation equality in the expand-ing branch. (Recall that matter-radiation equality oc-curs before recombination, and the superscript ‘+’ on t+edenotes the matter-radiation equality in the expandingpost-bounce branch.)

If we assume a symmetric bounce, then it follows thatthe times of matter-radiation equality before and afterthe bounce are symmetric around t = 0, in which casete = −6 × 1054tPl. From this and the relation (7), itis easy to check that kphys/|He| ∼ 1 is of the order ofunity rather than much smaller than 1. This shows that

3 Of course, this relation has to hold for all observed k, here wesimply choose a reasonable value of k in order to better under-stand the consequences of imposing (44).

8

Page 9: A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y 1Department of Physics, McGill University, Montr eal, QC H3A 2T8, Canada 2Department of Physics

- 3 x 1 0 1 2 - 2 x 1 0 1 2 - 1 x 1 0 1 2 0

1 0 5

1 0 6

1 0 7

η

λ = 1 / | |

λ s = 1 / c s k λ k = 1 / k

(a)

- 3 x 1 0 1 2 - 2 x 1 0 1 2 - 1 x 1 0 1 2 0

1 0 - 2 6

1 0 - 2 2

1 0 - 1 8

1 0 - 1 4

1 0 - 1 0

η

2

2

(b)

- 1 . 5 x 1 0 1 1 - 1 . 0 x 1 0 1 1 - 5 . 0 x 1 0 1 0 0 . 0 5 . 0 x 1 0 1 00 . 0 0

0 . 0 2

0 . 0 4

0 . 0 6

0 . 0 8

η

r = ∆

2 / ∆

2

(c)

FIG. 3. Evolutions of cosmological perturbations with a fixed comoving wave number k = 2.2 × 10−7 in the model under considerationin the frame of LQC. The left panel shows the comparison among the conformal Hubble radius λH = 1/H (the green solid line), thecomoving sound wavelength λε = 1/εk (the red dashed line), and the regular comoving wavelength λk = 1/k (the blue dotted line). Themiddle panel depicts the dynamics of primordial power spectra for curvature perturbations ∆2

R (the blue solid curve) and gravitationalwaves ∆2

h (the red dashed curve) along with the bouncing background. The right panel presents the evolution of the tensor-to-scalarratio r (the orange solid line). The initial conditions for the background field and model parameters are the same as for Fig. 2 and aregiven in (19). The initial conditions for cosmological perturbations of both scalar and tensor types are that they are initially quantumvacuum fluctuations in the early matter-dominated contracting phase.

a symmetric bounce is not viable in this model. To makethis conclusion explicit, we rewrite (39) as

a(t+e )H+e

a(te)He· a(to) k?(to)

a(t+e )H+e 1, (45)

which in turn, since the second term is of order unity andvia (5) and (7), gives

a(te)

a(t+e ) 1. (46)

This relation shows that the bounce must be significantlyasymmetric. Indeed, in order for the condition (44) tohold, |He| must be much larger (by at least a few or-ders of magnitude) than it would be in a model with asymmetric bounce, in which case it is necessary for thematter-radiation equality to occur at much higher curva-ture scales in the contracting branch than it does in theexpanding branch. One possible way for this to happenwould be if a very large number of additional quanta ofradiation are created during the bounce; as an aside notethat such a process would generate a significant amountof entropy.

Interestingly, it has recently been suggested that par-ticle production may play an important role during thebounce in LQC and this effect would cause the bounceto be significantly asymmetric in precisely the manneroutlined here [17]. That being said, it is not yet clearwhether particle production could generate the amountof asymmetry that is required for this realization of thematter bounce to be viable. We leave this question forfuture work.

There are also other reasons why an asymmetricbounce is necessary in this scenario. Assuming the Hub-ble parameter at te to be of the order of He ∼ 10−55t−1

Pl—as would be the case for a symmetric bounce— then

for the amplitude of the scalar perturbations to matchthe observed value of ∆2

R ∼ 10−9, it would be necessaryto have a very small value of the sound speed of cold darkmatter of cs = ε ∼ 10−15 (assuming ρc to be of approxi-mately the same order of magnitude as ρPl); such a smallvalue of the sound speed parameter typically leads to pri-mordial perturbations with over-large non-Gaussianities.In order to have a larger (although still small) value of εand thus sufficiently small non-Gaussianities, it is againnecessary to have an asymmetric bounce. This likely re-quires even more asymmetry than is needed to satisfy thecondition (46).

Finally, many of the modes observed today reenteredthe sound Hubble radius during radiation-domination.However, in order for them to be scale-invariant, in thecontracting branch of the universe, they must have ex-ited the sound Hubble radius during matter-domination.This is yet another reason that an asymmetric bounce isnecessary in the ΛCDM bounce scenario.

B. Numerical Analysis of the Perturbations

To complete the analysis of cosmological perturba-tions, in this subsection we perform a numerical compu-tation of the evolution of the primordial curvature per-turbations and gravitational waves in the model underconsideration. To be consistent with the background nu-merics, we consider the universe filled with a dust matterfield with its energy density evolving as (13), of which theinitial value is the same as the one given in (19). Further-more, we impose the initial conditions of the cosmologicalperturbations to be vacuum fluctuations during the mat-

9

Page 10: A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y 1Department of Physics, McGill University, Montr eal, QC H3A 2T8, Canada 2Department of Physics

ter dominated contracting phase (in units where ~ = 1),

Rinik → e−iεkη√2εk z

, hinik → e−ikη√2k a

. (47)

In addition, we take ε = 0.08 and fix k = 2.2 × 10−7 asan example in the detailed calculation. Our numericalresults are presented in Fig. 3.

In the left panel of Fig. 3, one can see how primordialcosmological perturbations evolve from the sub-Hubblescale to the super-Hubble region in the contracting phase.For a fixed comoving wave number k, the curvatureperturbation exits the Hubble radius much earlier thanthe gravitational wave since its oscillation gets squeezedwhen k ∼ H/cs at the sound horizon which is muchsmaller than the Hubble radius. Consequently, one ex-pects that there ought to be more oscillations in thepower spectrum of primordial gravitational waves thanthat of primordial curvature perturbations. This expec-tation is exactly verified in the middle panel of the figure,which displays the evolutions of primordial power spectraof both scalar and tensor perturbations.

In the middle part of Fig. 3, one can see that ∆2R ex-

periences only one oscillation and then becomes squeezedvery soon with its amplitude increasing until the nonsin-gular bounce takes place. However, ∆2

h experiences sev-eral oscillations during the contracting phase and onlybecomes squeezed when the universe is very near thebouncing phase. They both become conserved at super-Hubble scales after the bounce, which can be read fromthe right regime of Fig. 3b. Interestingly, one can ob-serve that the magnitudes of the two spectra are compa-rable (though the amplitude of the scalar mode is slightlylarger) during the contracting phase but the amplitudeof the scalar spectrum becomes significantly larger thanthat of tensor spectrum after the bounce, and hence thetensor-to-scalar ratio is suppressed to a small value thatis consistent with observations.

This can also be seen in the right panel of Fig. 3which depicts the dynamics of the tensor-to-scalar ra-tio r throughout the cosmological bouncing evolution. InFigs. 3b and 3c, we see that while the scalar perturbationmode passes through the bouncing phase in a relativelysmooth fashion, the magnitude of the gravitational wavemode is significantly damped. While the amplitude of thegravitational wave increases somewhat after the bounce,it remains significantly lower than before the bouncingphase.

In the specific example with ε = 0.08 being considered,we read approximately r ' 0.016 from the numericalcomputation. As we will analyze in the next section onthe tensor-to-scalar ratio, this is roughly in the same or-der of the analytical estimate (r ' 0.012) given by (54),though the detailed values are not in exact agreementwith each other. We will comment further on this slightnumerical discrepancy at the end of this section.

Another important point is that in the numerical com-putation we have chosen the critical density ρc to be

very small (ρc ∼ 10−9) so that the bounce scale is of or-der O(10−6) Planck mass (this choice for ρc significantlylowers the computational cost of the numerics as thiscauses the bounce to occur at a lower curvature scale).As a consequence, the amplitude of the power spectrumof primordial curvature perturbations is about O(10−12)which is approximately three orders lower than the mag-nitude of the observed CMB spectrum. The numericalcomputation performed in this subsection is sufficient todemonstrate the formation of primordial power spectrain the model under consideration and to verify their re-lations obtained in semi-analytic analyses. However, onecan improve the agreement of the amplitude of the spec-trum with observations by fine-tuning the values of theparameters in this model.

It is also important to keep in mind that it is possible toobtain a similar amplitude of scalar perturbations even ifρc ∼ ρPl, although such a choice would require a smallervalue of |He| and/or a larger value of ε.

Finally, note that since there are two matter compo-nents used in this step of the numerical computation,the sound speed parameter depends on the backgroundevolution approximately as

cs =

ε for Ωm ' 1√ω otherwise,

(48)

where ε is the small constant sound speed for CDM andω is the time-dependent equation of state ω = Ptot/ρtot,which includes the contributions coming from the radi-ation field and therefore equals 1/

√3 near the bounce

when radiation dominates the dynamics of the universe.This time-dependence effect has been taken into accountin the above numerical computation.

Finally, these numerical solutions also validate the ap-proximations made in the analytical section. First, we seethat in the near-bounce region it is justified to assumethat the modes of interest are in the long-wavelengthlimit. Furthermore, we have also numerically checkedthe difference between the solution obtained in the pre-vious section under the approximation of a discontinuouschange of the equation of state, and the result plotted inFig. 3b where no such approximation is made. We donot include the graph here, as the two curves lie practi-cally one on top of the other, showing the validity of thisapproximation. A careful comparison of the two curvesshows that the theoretical calculation given in Sec. II Avery slightly overestimates the amplitude of R.

This last point raises an important issue: the resultspresented in this paper rely either on analytic calcu-lations based on several approximations that are well-motivated but certainly introduce some small errors, oron numerical simulations that, while accurate enough forour purposes here, are not high-precision numerical stud-ies. Due to the use of these approximations and numer-ical simulations, the predictions presented in this papernecessarily contain some small errors. That being said,these small errors are not expected to affect the reliability

10

Page 11: A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y 1Department of Physics, McGill University, Montr eal, QC H3A 2T8, Canada 2Department of Physics

of the estimates obtained for the predicted observables inthe ΛCDM bounce scenario we consider here, which weestimate (by comparisons between the analytical and nu-merical results) to be accurate up to an overall factor ofapproximately 2.

IV. TENSOR PERTURBATIONS

It is possible to also calculate the spectrum of tensorperturbations after the bounce, again assuming that theinitial state was the quantum vacuum.

The LQC effective Mukhanov-Sasaki equation for ten-sor perturbations is4 [19]

µ′′ −(

1− 2ρ

ρc

)∇2µ− z′′T

zTµ = 0, (49)

where µ = h/zT , with h being the usual tensor pertur-bation modes, and

zT =a√

1− 2ρ/ρc. (50)

While the ‘time-dependent potentials’ z′′/z for scalarperturbations and z′′T /zT for tensor perturbations are notthe same, the most important difference for our purposesis the fact that the sound speed for tensor perturbationsis always 1, while the sound speed for scalar perturba-tions is significantly smaller than 1 during the matter-dominated phase. This effect strongly suppresses thetensor-to-scalar ratio, as we show below.

With the tensor power spectrum defined as

∆2h =

k3

2π264πG|h|2, (51)

and the tensor-to-scalar ratio r

r =∆2h

∆2R, (52)

it is possible to calculate the spectrum of the tensor per-turbations in a fashion analogous to Sec. III. While thecalculation is a little long, it is not particularly illumi-nating as it follows exactly the same steps as the one forscalar perturbations. While there are a few numerical

4 It is not possible to use the simplest separate universe modelsto handle tensor modes as they necessarily require off-diagonalelements in the metric, which do not appear in isotropic space-times. While it is likely that the separate universe approachcould be appropriately generalized by using the anisotropicBianchi cosmologies to model each ‘separate universe’, this hasnot been done yet. Instead, the equation of motion here is de-rived by demanding that the constraint algebra in the effectivetheory be anomaly-free. This procedure, when used to studyscalar perturbations, gives the same effective Mukhanov-Sasakiequation as what was obtained in lattice LQC [18].

factors that are different, the procedure is identical andtherefore here we will simply state the results.

The power spectrum of the tensor perturbations isfound to be almost scale-invariant, although the tilt de-pends on the value of the effective equation of state atthe time that the mode exits the Hubble radius. Recallthat the scalar modes reach the long wavelength limitwell before the tensor modes since, during the cold-dark-matter-dominated era, the sound speed for scalar modes

is c(S)s = ε while c

(T )s = 1 for the tensor modes. Therefore,

while the effective equation of state when the scalar per-turbations reach the long wavelength limit is ωeff = −δ,the effective equation of state when the tensor modes exitthe Hubble radius will have changed and will be larger(as the cosmological constant contributes less to the effec-tive equation of state for a smaller scale factor), thoughstill close to zero. We denote this value of the effectiveequation of state by −δT , and we can bound δT aboveby δT ≤ δ. Then the departure from scale-invariance isgiven by

nT = −12 δT . (53)

Note that δT , while expected to be small, may be neg-ative and this would give a slight blue tilt to the spec-trum of the tensor perturbations. Therefore, while near-scale-invariant tensor perturbations are predicted by thismodel, the exact departure from scale-invariance for thetensor modes will depend on how the effective equationof state varies in time, and this will determine whetherthere is a small blue tilt or a small red tilt.

Finally, for the particular scenario studied here, theamplitude of primordial gravitational waves is predictedto be very strongly suppressed, with a tensor-to-scalarratio of

r = 24 ε3, (54)

where ε refers to the sound speed of cold dark matterand r is therefore predicted to be very small. Note thatthe tensor-to-scalar ratio is suppressed both by a contri-bution due to the sound speed of cold dark matter, andalso by a further factor of 1/4 during the bounce due toquantum gravity effects.

There do not appear to be many estimates of the soundspeed of cold dark matter in the literature; one interest-ing reference is [20] which provides a bound of approx-imately ε2 . 0.03 (note however that in that paper theauthors study a considerably different model from thisone). It is likely that in the future a better estimate forε can be found, but nonetheless the constraint ε2 . 0.03already implies an upper bound of r . 0.12 on the tensor-to-scalar ratio, a result which is in agreement with thelatest observations [1, 2].

V. DISCUSSION

In this paper we have seen how in a contracting uni-verse cosmological perturbations, assumed to be initially

11

Page 12: A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y 1Department of Physics, McGill University, Montr eal, QC H3A 2T8, Canada 2Department of Physics

in their quantum vacuum state, become scale-invariant iftheir sound wavelength becomes larger than the Hubbleradius when the dynamics of the universe is dominated bycold dark matter. A small red tilt is generated when theeffective equation of state is negative due to the presenceof a positive cosmological constant, and a small tensor-to-scalar ratio is predicted.

The scale-invariant perturbations can provide appro-priate initial conditions for an expanding universe in or-der to seed structure formation if there is a bounce toconnect the contracting branch of the universe to our cur-rent expanding branch. In the realization of the matterbounce scenario studied here, the bounce occurs due tonon-perturbative quantum gravity effects, as captured byLQC, that resolve the classical singularity and provide aquantum bridge between the pre-big-bang and post-big-bang epochs. Further, the only matter fields present inthis model are assumed to be cold dark matter and radi-ation, together with a positive cosmological constant.

The main predictions depend on four parameters: thescalar index ns is determined by the effective equationof state at the time when the Fourier modes vk exit thesound Hubble radius, while the amplitude of the scalarperturbations depend on a combination of the soundspeed of cold dark matter ε, the amount of asymme-try in the universe, which can be parametrized by He

the Hubble rate at matter-radiation equality in the con-tracting branch, and the matter energy density at thebounce ρc. In particular, a symmetric bounce is ruledout and, in order to match observations, it is necessaryfor matter-radiation equality to occur at higher space-time curvature scales in the contracting branch than inthe expanding branch. This type of asymmetry can becaused by particle production during the bounce, a pro-cess which may be important in LQC [17].

There exist several other realizations of the matterbounce scenario, some where the bounce is caused bymatter fields that violate energy conditions [21–25], andothers where it is quantum gravity effects that providethe bounce [26–31]. There are also some matter bouncescenarios where the gravitational sector is modified notonly in the high-curvature regime, but also in the pre-bounce era in order to obtain a matter-like contractingphase without requiring the presence of any matter fields[32, 33]. Of course, there is no a priori reason to preferone realization of the matter bounce scenario over an-other (other than perhaps the simplicity or the eleganceof a particular model). Instead, it is necessary to de-termine how these realizations differ in their predictionswhich can then be compared to observations.

For this reason, it is important to point out that thepredictions of the different matter bounce realizationsvary in some significant aspects, concerning both CMBexperiments [34–38] as well as dark matter searches [39–41]. The recent review [38] explains the different predic-tions of many of the various realizations of the matterbounce scenario and in order to complement that paper,here we shall briefly explain how the predictions of the

model studied in this paper differ from the other realiza-tions of the matter bounce scenario.

First, the ΛCDM bounce model studied in this pa-per gives a slight red tilt in a natural fashion, somethingwhich is absent in many other realizations of the matterbounce (see for example [25, 28, 42] for specific realiza-tions). And second, the predicted tensor-to-scalar ratiois very small. Most realizations of the matter bounce typ-ically predict relatively large tensor-to-scalar ratios, andit is often necessary to assume the presence of a largenumber of fields (and thus of entropy perturbations aswell) in order to decrease the relative amplitude of thetensor perturbations. This is not necessary here since thetensor-to-scalar ratio is naturally predicted to be smalldue to the small sound speed of cold dark matter. Notethat this effect of a small cs strongly damping the valueof r has also previously been noticed in a study of thematter bounce scenario in a Bohm-de Broglie quantumcosmology model [43].

Another important point is that this scenario predictsa positive running of the scalar index, dns/dk > 0. Thisresult shows that it is possible for observations to dif-ferentiate between the ΛCDM bounce scenario and infla-tionary models, where the running of the scalar index ispredicted to be negative. The amplitude of this effect inthe ΛCDM bounce scenario depends on the time evolu-tion of effective equation of state during matter domina-tion −δ, which in turn depends quite sensitively on theparameters ΩΛ, Ωm and Ωr set as initial conditions forthe background in the contracting branch. We leave adetailed study of the relation between these parametersand the amplitude of dns/dk for future work.

There do remain two other important open questionsregarding this model that are also left for future work.The first one concerns the importance of particle pro-duction effects during the bounce. Will there be enoughparticle production to cause sufficient asymmetry aroundthe bounce point for this model to be viable? The sec-ond, more difficult, open problem is to determine how thepresence of anisotropies would modify the predictions ofthis model. The Friedmann equation for the mean scalefactor in Bianchi models shows that anisotropies dom-inate the dynamics in the high curvature limit, and weshould expect them to typically become important during(and possibly for some time before and after) the bouncein LQC. Indeed, anisotropies can modify the predictionsof a number of cosmological scenarios including inflation[44]. Despite their importance, anisotropies are oftenneglected as cosmological perturbation theory becomesconsiderably more complex in their presence [45–48]; inparticular, for non-vanishing anisotropies the equationsof motion for the scalar, vector and tensor perturbationsno longer decouple.

Note that if there is an ekpyrotic phase in the contract-ing branch of the space-time, it is possible to avoid thegrowth of anisotropies generated during matter contrac-tion [23, 24, 49–51] and then there is no need to includethem in the analysis. However, it is certainly possible

12

Page 13: A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y 1Department of Physics, McGill University, Montr eal, QC H3A 2T8, Canada 2Department of Physics

(and outside of ekpyrotic models it appears more natu-ral) for anisotropies to grow and dominate the dynamicsnear the bounce. In this case, while the anisotropies willbe diluted soon after the bounce (see, e.g., Sec. IIIC of[52]), they may change the spectrum of the cosmologi-cal perturbations as they evolve through the anisotropy-dominated bounce, and some of these modifications mayultimately be observable today. For this reason, it isimportant to allow for anisotropies by generalizing theresults obtained here for the case where the backgroundis an anisotropic Bianchi I space-time (rather than flatFLRW), and determining precisely how anisotropies mayaffect the predictions of this model.

Finally, we conclude with some comments regardingthe universality and robustness of the predictions of theΛCDM bounce scenario with respect to the physics ofthe bounce. In this paper we assumed that the bounce iscaused by the quantum gravity effects captured by loopquantum cosmology. As mentioned above, it is also pos-sible to generate a bounce via other mechanisms, eitherwith matter fields that violate energy conditions, or bymodifications to the gravitational action as occur for ex-ample in f(R) theories. An important property of theΛCDM bounce scenario is that most of the salient char-acteristics of the predicted spectra of scalar and tensorperturbations are due to the pre-bounce physics, wherequantum gravity and other high energy effects are neg-ligible and therefore the majority of the predictions —

including the small red tilt for the modes that exit thesound Hubble radius when the effective equation of stateis slightly negative, as well as the positive running of thescalar index— are independent of the high-curvature dy-namics of the background space-time at the bounce point.However, there is one effect that is partially due to LQC:the tensor-to-scalar ratio is suppressed by a factor of 1/4during the bounce by LQC effects (specifically, due to themodifications that arise in the LQC effective Mukhanov-Sasaki equation for tensor perturbations). Thus, whilemany predictions of the ΛCDM bounce are quite robustand are mostly independent of the physics of the bounce,one exception is the predicted value of the tensor-to-scalar ratio which in fact is affected by LQC effects and,if the sound speed of cold dark matter is known, this ef-fect could provide a potentially important observationaltest for LQC.

ACKNOWLEDGMENTS

We would like to thank Robert Brandenberger, JimCline, Parampreet Singh, Jun-Qing Xia and Wei Xue forhelpful discussions. The work of YFC is supported inpart by an NSERC Discovery grant and by funds from theCanada Research Chair program. This work is supportedin part by a grant from the John Templeton Foundation.

[1] WMAP, G. Hinshaw et al., “Nine-Year WilkinsonMicrowave Anisotropy Probe (WMAP) Observations:Cosmological Parameter Results,” Astrophys. J. Suppl.208 (2013) 19, arXiv:1212.5226.

[2] Planck Collaboration, P. Ade et al., “Planck 2013results. XVI. Cosmological parameters,” Astron.Astrophys. (2014) arXiv:1303.5076.

[3] M. Bojowald, “Loop quantum cosmology,” Living Rev.Rel. 11 (2008) 4.

[4] A. Ashtekar and P. Singh, “Loop Quantum Cosmology:A Status Report,” Class. Quant. Grav. 28 (2011)213001, arXiv:1108.0893.

[5] K. Banerjee, G. Calcagni, and M. Martın-Benito,“Introduction to loop quantum cosmology,” SIGMA 8(2012) 016, arXiv:1109.6801.

[6] D. Wands, “Duality invariance of cosmologicalperturbation spectra,” Phys. Rev. D60 (1999) 023507,arXiv:gr-qc/9809062.

[7] F. Finelli and R. Brandenberger, “On the generation ofa scale invariant spectrum of adiabatic fluctuations incosmological models with a contracting phase,” Phys.Rev. D65 (2002) 103522, arXiv:hep-th/0112249.

[8] V. Taveras, “Corrections to the Friedmann Equationsfrom LQG for a Universe with a Free Scalar Field,”Phys. Rev. D78 (2008) 064072, arXiv:0807.3325.

[9] C. Rovelli and E. Wilson-Ewing, “Why are the effectiveequations of loop quantum cosmology so accurate?,”Phys. Rev. D90 (2014) 023538, arXiv:1310.8654.

[10] P. Singh and V. Taveras, “A note on the effectiveequations in LQC,” unpublished.

[11] T. Paw lowski, R. Pierini, and E. Wilson-Ewing, “Loopquantum cosmology of a radiation-dominated flatFLRW universe,” arXiv:1404.4036.

[12] V. F. Mukhanov, H. Feldman, and R. H.Brandenberger, “Theory of cosmological perturbations,”Phys. Rept. 215 (1992) 203–333.

[13] D. Salopek and J. Bond, “Nonlinear evolution of longwavelength metric fluctuations in inflationary models,”Phys. Rev. D42 (1990) 3936–3962.

[14] D. Wands, K. A. Malik, D. H. Lyth, and A. R. Liddle,“A New approach to the evolution of cosmologicalperturbations on large scales,” Phys. Rev. D62 (2000)043527, arXiv:astro-ph/0003278.

[15] E. Wilson-Ewing, “Holonomy Corrections in theEffective Equations for Scalar Mode Perturbations inLoop Quantum Cosmology,” Class. Quant. Grav. 29(2012) 085005, arXiv:1108.6265.

[16] E. Wilson-Ewing, “Lattice loop quantum cosmology:scalar perturbations,” Class. Quant. Grav. 29 (2012)215013, arXiv:1205.3370.

[17] A. T. Mithani and A. Vilenkin, “Instability of anemergent universe,” JCAP 1405 (2014) 006,arXiv:1403.0818.

[18] T. Cailleteau, J. Mielczarek, A. Barrau, and J. Grain,“Anomaly-free scalar perturbations with holonomycorrections in loop quantum cosmology,” Class. Quant.Grav. 29 (2012) 095010, arXiv:1111.3535.

13

Page 14: A CDM Bounce Scenario · A CDM Bounce Scenario Yi-Fu Cai1, and Edward Wilson-Ewing2,3, y 1Department of Physics, McGill University, Montr eal, QC H3A 2T8, Canada 2Department of Physics

[19] T. Cailleteau, A. Barrau, J. Grain, and F. Vidotto,“Consistency of holonomy-corrected scalar, vector andtensor perturbations in Loop Quantum Cosmology,”Phys. Rev. D86 (2012) 087301, arXiv:1206.6736.

[20] A. Balbi, M. Bruni, and C. Quercellini, “Lambda-alphaDM: Observational constraints on unified dark matterwith constant speed of sound,” Phys. Rev. D76 (2007)103519, arXiv:astro-ph/0702423.

[21] Y.-F. Cai, T.-T. Qiu, R. Brandenberger, and X.-M.Zhang, “A Nonsingular Cosmology with aScale-Invariant Spectrum of Cosmological Perturbationsfrom Lee-Wick Theory,” Phys. Rev. D80 (2009)023511, arXiv:0810.4677.

[22] C. Lin, R. H. Brandenberger, and L. PerreaultLevasseur, “A Matter Bounce By Means of GhostCondensation,” JCAP 1104 (2011) 019,arXiv:1007.2654.

[23] Y.-F. Cai, D. A. Easson, and R. Brandenberger,“Towards a Nonsingular Bouncing Cosmology,” JCAP1208 (2012) 020, arXiv:1206.2382.

[24] Y.-F. Cai, E. McDonough, F. Duplessis, and R. H.Brandenberger, “Two Field Matter BounceCosmology,” JCAP 1310 (2013) 024, arXiv:1305.5259.

[25] Y.-F. Cai, “Exploring Bouncing Cosmologies withCosmological Surveys,” Sci. China Phys. Mech. Astron.57 (2014) 1414–1430, arXiv:1405.1369.

[26] T. Biswas, A. Mazumdar, and W. Siegel, “Bouncinguniverses in string-inspired gravity,” JCAP 0603 (2006)009, arXiv:hep-th/0508194.

[27] R. Brandenberger, “Matter Bounce in Horava-LifshitzCosmology,” Phys. Rev. D80 (2009) 043516,arXiv:0904.2835.

[28] E. Wilson-Ewing, “The Matter Bounce Scenario inLoop Quantum Cosmology,” JCAP 1303 (2013) 026,arXiv:1211.6269.

[29] R. H. Brandenberger, C. Kounnas, H. Partouche, S. P.Patil, and N. Toumbas, “Cosmological PerturbationsAcross an S-brane,” JCAP 1403 (2014) 015,arXiv:1312.2524.

[30] P. Peter, E. J. Pinho, and N. Pinto-Neto, “A Noninflationary model with scale invariant cosmologicalperturbations,” Phys. Rev. D75 (2007) 023516,arXiv:hep-th/0610205.

[31] S. Odintsov and V. Oikonomou, “Matter Bounce LoopQuantum Cosmology from F (R) Gravity,”arXiv:1410.8183.

[32] G. Leon and A. A. Roque, “Qualitative analysis ofKantowski-Sachs metric in a generic class of f(R)models,” JCAP 1405 (2014) 032, arXiv:1308.5921.

[33] K. Bamba, A. N. Makarenko, A. N. Myagky, and S. D.Odintsov, “Bouncing cosmology in modifiedGauss-Bonnet gravity,” Phys. Lett. B732 (2014)349–355, arXiv:1403.3242.

[34] Y.-F. Cai, T.-T. Qiu, J.-Q. Xia, and X. Zhang, “AModel Of Inflationary Cosmology Without Singularity,”Phys. Rev. D79 (2009) 021303, arXiv:0808.0819.

[35] Y.-F. Cai and X. Zhang, “Primordial perturbation witha modified dispersion relation,” Phys. Rev. D80 (2009)

043520, arXiv:0906.3341.[36] J. Liu, Y.-F. Cai, and H. Li, “Evidences for bouncing

evolution before inflation in cosmological surveys,” J.Theor. Phys. 1 (2012) 1–10, arXiv:1009.3372.

[37] J.-Q. Xia, Y.-F. Cai, H. Li, and X. Zhang, “Evidencefor bouncing evolution before inflation after BICEP2,”Phys. Rev. Lett. 112 (2014) 251301, arXiv:1403.7623.

[38] Y.-F. Cai, J. Quintin, E. N. Saridakis, andE. Wilson-Ewing, “Nonsingular bouncing cosmologies inlight of BICEP2,” JCAP 1407 (2014) 033,arXiv:1404.4364.

[39] C. Li, R. H. Brandenberger, and Y.-K. E. Cheung, “BigBounce Genesis,” arXiv:1403.5625.

[40] Y.-K. E. Cheung, J. U. Kang, and C. Li, “Dark matterin a bouncing universe,” JCAP 1411 (2014) 001,arXiv:1408.4387.

[41] Y.-K. E. Cheung and J. Vergados, “Direct dark mattersearches - Test of the Big Bounce Cosmology,”arXiv:1410.5710.

[42] E. Elizalde, J. Haro, and S. D. Odintsov, “Quasi-matterdomination parameters in bouncing cosmologies,”arXiv:1411.3475.

[43] D. Bessada, N. Pinto-Neto, B. B. Siffert, and O. D.Miranda, “Stochastic background of relic gravitons in abouncing quantum cosmological model,” JCAP 1211(2012) 054, arXiv:1207.5863.

[44] C. Pitrou, T. S. Pereira, and J.-P. Uzan, “Predictionsfrom an anisotropic inflationary era,” JCAP 0804(2008) 004, arXiv:0801.3596.

[45] M. Den, “Gauge Invariant Perturbations in a SpatiallyFlat Anisotropic Universe,” Prog. Theor. Phys. 77(1987) 653.

[46] H. Noh and J. Hwang, “Perturbations of an anisotropicspace-time: Formulation,” Phys. Rev. D52 (1995)1970–1987.

[47] P. K. Dunsby, “Gauge invariant perturbations ofanisotropic cosmological models,” Phys. Rev. D48(1993) 3562–3576.

[48] T. S. Pereira, C. Pitrou, and J.-P. Uzan, “Theory ofcosmological perturbations in an anisotropic universe,”JCAP 0709 (2007) 006, arXiv:0707.0736.

[49] Y.-F. Cai, R. Brandenberger, and P. Peter, “Anisotropyin a Nonsingular Bounce,” Class. Quant. Grav. 30(2013) 075019, arXiv:1301.4703.

[50] Y.-F. Cai and E. Wilson-Ewing, “Non-singular bouncescenarios in loop quantum cosmology and the effectivefield description,” JCAP 1403 (2014) 026,arXiv:1402.3009.

[51] S. Alexander, Y.-F. Cai, and A. Marciano,“Fermi-bounce cosmology and the fermion curvatonmechanism,” arXiv:1406.1456.

[52] B. Gupt and P. Singh, “Quantum gravitational Kasnertransitions in Bianchi-I spacetime,” Phys. Rev. D86(2012) 024034, arXiv:1205.6763.

14