Bouncing cosmology in a curved braneworld
Abstract
We explore the possibility of a non-singular bounce in our universe from a warped braneworld scenario with dynamical branes and a non-zero brane cosmological constant. Such models naturally incorporate a scalar sector known as the radion originating from the modulus of the theory. The presence of brane cosmological constant renders the branes to be non-flat and gives rise to a potential and a non-canonical kinetic term for the radion field in the four dimensional effective action. The kinetic term exhibits a phantom-like behavior within the domain of evolution of the modulus which leads to a violation of the null-energy condition often observed in a bouncing universe. The interplay of the radion potential and kinetic term enables the evolution of the radion field from a normal to a phantom regime where the universe transits from a contracting era to an expanding epoch through a non-singular bounce. Analysis of the scalar and tensor perturbations over such background evolution reveal that the primordial observables e.g., the amplitude of scalar perturbations , tensor to scalar ratio and the scalar spectral index are in agreement with the current constraints reported by the Planck satellite. The implications are discussed.
1 Introduction
One of the major challenges in modern theoretical cosmology is to explain the early stage of the universe, in particular, whether the universe emerged from an initial singularity (also known as the Big-Bang singularity) or the universe underwent a non-singular bounce leading to a possible singularity free expansion of the universe. Some of the early universe scenarios, proposed so far, that can generate an almost scale invariant power spectrum and hence confront the observational constraints are the inflationary scenario [1, 2, 3, 4, 5, 6, 7, 8], the bouncing universe [9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53, 54, 55, 58, 59, 60, 61, 62, 63, 64, 65, 66, 67, 70, 56, 68, 69, 57], the emergent universe scenario [71, 72, 73, 74, 75, 76] and the string gas cosmology [77, 78, 79, 80, 81, 82, 83].
In this work, we study the bouncing scenario in a non-flat warped braneworld model. The bouncing scenario consists of two erasan era of contraction and an era of expansion of the scale factor, both the eras being connected by a non-singular bounce [9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53, 54, 55, 58, 59, 60, 61, 62, 63, 64, 65, 66, 67, 70, 56, 68, 69, 57]. Beside producing an observationally compatible primordial power spectrum, the bouncing scenario has the merit to give rise to a singularity free evolution of the early universe. Although it has been argued that the Big-Bang singularity may be avoided through a suitable quantum generalization of gravity, the absence of a consistent quantum theory of gravity makes the bouncing description of the universe a promising scenario. In this context it may be metioned that string theory [86, 87, 84, 85] inherently incorporates the quantum nature of gravity and is associated with several extra spatial dimensions. Although originally intended to unify the known forces of nature, it turns out that extra dimensions can also provide plausible resolution to the gauge-hierarchy problem or the finetuning problem in particle physics arising due to large quantum corrections of the Higgs mass [88, 86, 87, 89, 90, 91, 92, 93, 94, 95, 96, 97]. In this context warped geometry models are particularly relevant. In particular, the warped geometry models due to Randall-Sundrum (RS) [94] earned a lot of attention since it resolves the gauge hierarchy problem without introducing any intermediate scale (between Planck and TeV scale) in the theory.
The RS scenario consists of an extra spatial dimension (over the usual four dimensional spacetime) with orbifold symmetry, enclosed between two 3-branes which are considered to be flat. The distance between the two branes governs the magnitude of the brane warping and therefore plays the crucial role in resolving the gauge-hierarchy problem. The assumption of flat branes which gives rise to a vanishing brane cosmological constant in the RS scenario can be relaxed in a generalized warped braneworld model [98], which allows the branes to be non-flat giving rise to de Sitter (dS) or anti-de Sitter (AdS) branes. The interplay of the brane warping (which depends on the interbrane distance ) and the magnitude of the brane cosmological constant leads to the resolution of the gauge-hierarchy problem in such models. The cosmological, astrophysical and phenomenological implications of warped braneworld models (with flat or non-flat branes) have been discussed in [99, 100, 101, 102, 103, 104, 105, 106, 107, 108, 109, 110, 111, 112, 113, 114, 115, 116, 117, 118, 119, 120, 121]. Since the resolution of the finetuning problem essentially depends on the interbrane distance , the stabilization of to the appropriate value becomes crucial. The stabilization is achieved in the RS scenario by introducing a scalar field in the bulk [123, 124], which leads to a potential for in the four dimensional effective action whose minima can be suitably adjusted to address the finetuning problem. The origin of the bulk scalar however remains unexplored. This problem can evaded in the non-flat warped braneworld scenario with dynamical branes such that the interbrane distance attains the status of a field (the so called radion or the modulus). Such a framework enables the radion to generate its own potential along with a non-canonical kinetic term in the four dimensional effective action which in turn can stabilize the modulus to the suitable value [125, 120], without invoking any additional scalar field in the theory.
The non-canonical scalar kinetic term becomes negative for certain values of the modulus which endows the radion a phantom-like behavior where the null energy condition is violated. Such a violation is a generic feature observed in a bouncing universe, which motivates us to explore the prospect of the non-flat warped braneworld model in addressing bouncing cosmology. We investigate the cosmological evolution of the radion field in the FRW background and the subsequent evolution of the primordial fluctuations which allows us to understand the viability of the model in purview of the Planck 2018 constraints.
The paper is organized as follows: in 2, we briefly describe the non-flat warped braneworld model and its four dimensional effective theory. Having set the stage, 3 is dedicated for studying the background cosmological evolution while the evolution of the perturbations and confrontation of the theoretical predictions with the latest Planck observations is discussed in 4. We conclude with a summary of our results and a discussion of our findings in 5.
2 The non-flat warped braneworld scenario
Randall & Sundrum (RS)[94] proposed the warped braneworld scenario to address the fine-tuning problem in particle physics. The RS model consists of a 5-dimensional AdS bulk bounded by two 3-branes, namely the visible brane (where our 4-d universe resides) and the hidden brane. The extra dimension denoted by is associated with a orbifold symmetry and the hidden brane resides at while the visible brane is located at . In the RS scenario, the bulk metric is described by
(1) |
from which it is evident that the branes are flat which is ensured by the exact cancellation of the brane tension and the cosmological
constant induced on the brane [126].
In 1 the warp factor is represented by with where is the compactification radius and
such that and denote the five dimensional cosmological constant and Planck mass respectively.
The presence of the exponential warp factor in the metric ensures that is sufficient to bring down the Higgs mass from the
Planck scale to the TeV scale on the visible brane without introducing any new energy scale in the theory.
A generalization of the RS model to incorporate non-flat branes is important since the non-flatness of our universe is often evident from the physical situations, e.g., an expanding universe, the presence of black holes etc. This is achieved by replacing the brane metric with in the above metric ansatz. This induces a non-zero cosmological constant on the brane inherited from the bulk which can be both positive or negative. In the situation where the branes are de-Sitter, the warp factor is given by [98]:
(2) |
where, is a dimensionless constant directly proportional to the brane cosmological constant and . It can be shown that the above warp factor can give rise to the requisite warping of the Higgs mass on the visible brane (i.e., ) as in the RS scenario while keeping (the magnitude of the present day cosmological constant in Planckian units).
For AdS branes, it can be shown that the warp factor is given by , with [98]. We further note that in the event , we retrieve the RS warp factor describing the flat braneworld scenario, for both the dS and AdS branes. Since the observed accelerated expansion of the universe can be explained by a positive brane cosmological constant, we will concentrate mainly on the warped braneworld scenario with de-Sitter branes and investigate its role in bouncing cosmology.
2.1 The non-flat warped braneworld with the radion field
In the warped braneworld scenario, the resolution of the gauge hierarchy problem depends crucially on , the stable distance between the two branes. This requires a mechanism to stabilize the inter-brane distance to the suitable value. Goldberger & Wise [123] addressed this by invoking a bulk scalar in the five dimensional action which resulted in a potential for in the 4-dimensional effective action. They showed that the stable value of corresponds to the minima of the potential. However, the physical origin of the scalar field in the bulk action is not well understood.
Instead of flat branes if one considers the non-flat warped braneworld scenario, and allows the inter-brane distance to be treated as a 4-dimensional field (the so called radion or the modulus), then it can be shown that a potential for the modulus is naturally generated in the 4-d effective action which in turn can stabilize the modulus [125]. This modulus potential is completely attributed to the non-flat character of the branes and in the event the branes are flat this potential identically vanishes.
This scenario is described by the bulk action,
(3) |
such that,
(4) |
(5) |
(6) |
where is the bulk Ricci scalar and the determinant of the bulk metric . In 5, and refer to the brane tension and matter Lagrangian on the visible brane while and in 6 correspond to the brane tension and matter Lagrangian on the hidden brane .
The bulk is governed by the Einstein’s equations with the following solution for the metric,
(7) |
which is easily extended from 1 with replaced by and substituted by the radion field . We will concentrate on de-Sitter branes in this work and hence the form of the warp factor is given by
(8) |
which is the same as 2 with replaced by .
It is interesting to note that the the positivity of the warp factor in 8 requires that which immediately follows when we write the warp factor in the following way,
(9) |
and demand . Further, since the modulus cannot be negative the maximum value that can attain is unity, when . Therefore, throughout this work our region of interest in the field space would be . We will use this property of the warp factor when we explore bouncing cosmolgy with the radion field in 3.
The effective action in four dimensions is derived from the bulk action by integrating over the extra coordinate . This can be segregated into three parts, namely,
(10) |
where,
(11) |
is the curvature dependent part of the effective action with the determinant and the Ricci scalar with respect to the brane metric . From 11 it is evident that the Ricci scalar involves a coupling with the dimensionless radion field with and hence is in the Jordan frame. Here and in the rest of the discussion we shall denote as the radion field. The coupling of the modulus to is denoted by which assumes the form,
(12) |
The second part of the effective action comprises of a potential for the radion field,
(13) |
with
(14) |
It is interesting to note that the potential is directly proportional to and vanishes in the event the branes are flat i.e., [124]. Moreover, it has an inflection point at which will have important consequences when we explore early universe cosmology in this model.
The third part of the effective action in 10 is associated with the kinetic term of the radion given by,
(15) |
where,
(16) |
denotes the non-canonical coupling to the kinetic term which reduces to the canonical form when . Therefore, the non-flatness of the branes generates the brane cosmological constant which in turn gives rise to the potential for the radion and its non-canonical kinetic term.
Since the observations are generally made in the Einstein frame, we perform a conformal transformation of the Jordan frame metric to remove the coupling of the scalar field to the Ricci scalar. This is achieved by scaling the Jordan frame metric with the conformal field such that the metric in the Einstein frame is given by . With this conformal scaling it can be shown that in four dimensions, the Ricci scalar in the Einstein frame is related to the Ricci scalar in the Jordan frame by,
(17) |
where denotes covariant derivative with respect to the metric . Using 17 and the fact that and choosing , we arrive at the effective action in the Einstein frame,
(18) |
where and the potential due to the radion field in the Einstein grame is given by,
(19) |
while the non-canonical coupling to the kinetic term is given by,
(20) |
where ‘prime’ here implies differentiation with respect to . In 1a and 1b we plot the variation of and with the radion field for . It can be shown from 19 that the radion potential in the Einstein frame continues to have an inflection point at which can be confirmed from the vanishing first and second derivatives but a positive third derivative of with respect to at . Moreover, 1b reveals that the non-canonical coupling to the kinetic term exhibits a transition from a normal to a phantom regime (i.e from to ), where the phantom like behavior remains when lies in the range , with denoting the zero crossing of . Also note from 1b that for , .
We thus note that in the non-flat warped braneworld scenario, we have a scalar field, the radion, which is associated with a potential and a non-canonical kinetic term. It is believed that in the early universe the big bang singularity can be avoided in a bouncing scenario which is triggered by a scalar field with a potential. This raises the question whether the radion field can be instrumental in giving rise to a bouncing universe which we address in the next section.


3 Implications in Early Universe Cosmology: Background evolution
In this section we explore the role of the radion field in triggering a bouncing universe which can potentially avoid the big bang singularity. The bouncing scenario often invokes a scalar field with a potential and there exist plenty of models in the literature which can give rise to such a scenario (see [11]). In most of the cases the scalar potentials are reconstructed to explain the observations and often their origin remains unexplained. The merit of the non-flat warped braneworld model lies in the fact that the the radion field naturally arises from compactification in the effective four-dimensional theory and generates its own potential and non-canonical kinetic term. Here we consider the implications of the radion field in inducing a bouncing universe.
Since we are interested to study early universe cosmology with the radion field we consider the metric in the Einstein frame to be described by the FRW spacetime in the spatially flat form,
(21) |
with is known as the scale factor of the universe. In the generalized RS scenario, the 3-branes can be Minkowskian, de-Sitter or anti de-Sitter depending on the values of the induced brane cosmological constant [98]. Recall, in the present work, we consider the branes to be de-Sitter, i.e our visible universe is a 3-brane (i.e having three spatial dimension along with the time coordinate), described by the spatially flat FRW metric. Since the FRW metric is curved irrespective of its spatial curvature, the visible 3-brane is non-flat. At this stage it deserves mentioning that the spatially flat FRW metric is more consistent over the closed or open FRW universe from latest Planck 2018 data through TT, TE, EE + lowE + lensing + BAO data, where TT means temperature temperature cross-correlation of CMB data, TE means cross-correlation between temperature and electric type polarization of CMB data and finally BAO stands for Baryon Acoustic Oscillation [122]. Due to the time dependency of the scale factor, the metric in 21 indicates a non-zero curvature on the four dimensional brane geometry and moreover the brane curvature is characterized by the corresponding Ricci scalar given by . As we will see from 25 that the kinetic as well as the potential energy of radion field contributes to the on-brane Ricci scalar via the effective four dimensional Freidmann equation. Furthermore, as shown in 19, the potential energy of the radion field is proportional to the induced brane cosmological constant and thus one may argue that the radion potential energy is generated entirely due to the presence of the non-zero brane cosmological constant. Thereby the brane cosmological constant affects the evolution of the scale factor through the potential energy density of the radion field. Below, we will show that the metric ansatz of 21 is consistent with the field equations of motion and moreover it will lead to a non-singular bounce on our visible brane.
It is evident from 18 that the energy momentum tensor due to the radion field is given by,
(22) |
such that
(23) |
represents the energy density while
(24) |
corresponds to the pressure due to the radion field. We note that the radion field depends only on time since the background metric given by 21 is only time dependent.
Using 23 the Friedmann equation obtained from the temporal component of the Einstein’s equations assume the form,
(25) |
while the Friedman equation derived from the spatial component of the Einstein’s equations is given by,
(26) |
where denotes the Hubble parameter. The equation of motion for the radion field is given by,
(27) |
Using 23 and 24, 27 can be written as,
(28) |
25, 26 and 28 are the background equations, although it is important to note that 28
is not independent but can be derived from 25 and 26.
At this stage it deserves mentioning that in a canonical scalar tensor theory, the
Friedmann equation becomes (26) and thus the Hubble parameter decreases monotonically with cosmic time. Therefore a bounce phenomena is impossible in a canonical scalar tensor model as it cannot give rise to which is one of the necessary conditions
to get a bounce. On the contrary, in a non-canonical scalar tensor theory where the scalar field has non-canonical kinetic term (as in
the present context), the Friedmann equations are modified due to the presence of and the modified equations are given by
25 and 26 respectively. 26 clearly indicates that in a non-canonical scalar tensor model, the sign of actually
controls the energy condition, in particular leads to a violation of null energy condition which in turn may ensure a bouncing
phase in our visible universe. We have already noted in 3 that in the present scalar-tensor model the non-canonical kinetic term exhibits a transition
from a normal to a phantom regime (i.e from to ) where the null energy condition is violated. Therefore it is important to investigate the prospect of bouncing cosmology with the present non-flat warped braneworld model which we explore next. In particular, we first present the background evolution of and (governed by 25 and 26)
in the next section and subsequently study the evolution of the perturbations in 4.
In general, a non-singular bounce is characterized by the conditions and where is the cosmic time when
the bounce occurs. Keeping these conditions in mind, if we look into 25 and 26, then it is evident that the model has a possibility to show
a bounce phenomena when the non-canonical function becomes negative i.e when the radion field is in the phantom regime. The analysis in
3 reveals that is indeed negative in the regime . Thus, at first we analytically
solve the background equations near to investigate the bounce and then we numerically determine
the background evolution for a wide range of (or equivalently for a wide range
of cosmic time), where the boundary conditions of the numerical calculation are provided from the previously found analytic solutions.
In particular we consider,
(29) |
with . Due to the above form of , in 12 simplifies to
(30) |
such that is given by,
(31) |
while can be approximated as,
(32) |
The above simplifications in the form of and hold only in the regime where is given by 29 with . With these simplifications the evolution equations for the Hubble parameter and the radion field (i.e 25 and 26) turn out to be,
(33) |
and
(34) |
respectively. 33 can be solved to obtain the time evolution of the Hubble parameter which assumes the form,
(35) |
such that is given by,
(36) |
(37) |
with and solving the above equation yields the following time evolution for ,
(38) |
where is an integration constant which can be determined by demanding,
(39) |
which implies that (i.e the radion field) monotonically decreases with time and asymptotically goes to the value which is the minimum possible value of from the requirement of positive warp factor, as discussed after 9. From the condition given by 39, the form of turns out to be,
(40) |
which when substituted in 38 gives,
(41) |
With this, we arrive at the background solution for and in the regime with ,
(42) |
where, . 42 clearly indicates and at (corresponding to the bounce time) which are the necessary conditions for a non-singular bounce. Therefore, the present non-flat warped braneworld model predicts a bouncing universe in the visible brane when the radion field lies within the phantom regime, in particular near . In the phantom regime, due to the negative kinetic energy of the radion field, the effective null energy condition (NEC) is violated and makes the bounce possible at a certain finite time, in particular at . Here we would like to mention that such NEC violation occurs irrespective of any values of and (i.e the model parameters), and moreover the initial conditions of the solution of 42 is free from fine tuning of the model parameters. Thereby, we may argue that the bounce solution in the present context is a generic feature and does not require any fine-tuned values of the model parameters.
At this stage, it is important to check
whether the radion field, starting from a value in the normal regime, will reach to the phantom regime by its . For this purpose,
we solve the coupled equations for and (i.e 25 and 26) for a wide range of cosmic time numerically.
In regard to the numerical calculation, the boundary conditions are provided
from the analytic solutions as determined in 42, in particular the boundary conditions are given by and
, where we consider
(later, during the perturbation calculation, we show that such a value of is consistent with
the Planck 2018 constraints). The time evolution of the Hubble parameter and the radion field are shown in 2a and 2b respectively
(the radion field plot is magnified 1000 times i.e ).
In the inset of 2b,
the magenta curve denotes the time evolution of while the blue curve represents the zoomed-in version of
near the zero crossing of .
From 2b it is evident that exhibits a transition from a normal regime
(where ) to a phantom regime (where )
with its zero crossing occurs at a finite time before the bounce at . Moreover 2b demonstrates that
there is no divergence in the dynamical evolution of as transits from normal to the phantom regime.
However on the other hand, as evident from 26,
diverges at the time when the non-canonical kinetic coupling makes the zero crossing. Here we would like to mention that such divergence of
does not lead to any pathology to the radion field equation of motion i.e to 28 and the reason is following: 28 can be
equivalently expressed as which
includes and its derivative with respect to cosmic time. Now 26 evidents that is proportional to
which, along with its derivative, is indeed finite for all possible cosmic time (see 2a).
Thereby the term and its derivative with respect to are finite everywhere even at ,
and thus we may argue that the radion field equation of motion does not lead to any inconsistency in the present context.


2a reveals that the Hubble parameter becomes zero and increases with respect to cosmic time at , which confirms a non-singular bounce at . Before demonstrating the dynamics of the radion field, we recall that for , the zero crossing of occurs at , as shown in 1b i.e. exhibits the normal to phantom transition as crosses from higher values. Numerical solution of 25 and 26 indicates that the radion field starts its journey from the normal regime (i.e ) and dynamically moves to the phantom era (i.e ) with time by monotonically decreasing in magnitude and asymptotically stabilizes to the value which for . This is in accordance with the analytical results obtained in 42 and the time evolution of the background radion field is explicitly illustrated in 2b. As the radion field asymptotically tends to (i.e., ), the warp factor which in turn resolves the gauge-hierarchy problem for while the stabilized inter-brane separation [125]. Therefore, in the non-flat warped braneworld scenario with dynamical branes, the radion generates its own potential which in turn stabilizes the modulus dynamically in the FRW background. Further, the presence of the phantom era enables violation of the null energy condition for the radion field which makes this a promising model to explore the bouncing scenaio.
At this stage, it may be mentioned that a holonomy improved non-canonical scalar tensor model may rescue the energy condition in a bouncing scenario. In the holonomy generalized model, the squared Hubble parameter (i.e in 25) is proportional to the linear as well as quadratic power of energy density, unlike the usual Friedmann equations where is proportional only to the linear power of energy density. Such difference in the field equations may play a significant role to rescue the null energy condition necessary for a non-singular bounce. This investigation is expected to be carried out soon in a future work.
4 Implications in Early Universe Cosmology: Evolution of perturbations
In this section, we consider the spacetime perturbations over the background FRW metric and consequently determine the primordial observable quantities like the scalar spectral index (), tensor to scalar ratio () and the amplitude of scalar perturbations (). In a bouncing universe, the Hubble parameter becomes zero and consequently the comoving Hubble radius diverges at the bouncing point. However, the asymptotic behaviour of the Hubble radius differentiates various bouncing models which can be broadly classified into two scenarios. In the first case, the comoving Hubble radius decreases and goes to zero asymptotically with time, which corresponds to a late time accelerating universe. In this case, the perturbation modes generate near the bounce, because at that time, the horizon has an infinite size and all the perturbation modes lie within the horizon. In the second situation the Hubble radius diverges asymptotically with time, which indicates a decelerating universe at late time and consequently the primordial perturbation modes relevant for the present era generate at a distant past far away from the bounce. More explicitly, in the latter case, the comoving wave number begins its journey from the infinite past in the contracting universe, within the sub-Hubble scale, exits the horizon as it contracts, and again re-enters the horizon in the low curvature regime of the expanding phase and becomes relevant for present time observations. Therefore, depending on the asymptotic behavior of the Hubble radius, the perturbation modes in a bounce model generate either near the bounce or far away from the bounce deeply in the contracting regime.
Based on the above arguments, before moving to the perturbation calculations, we would like to investigate the asymptotic behaviour of the comoving Hubble radius (defined by ) in the context of present model. Using the background solution of the Hubble parameter from 2a, we give the evolution of with respect to cosmic time in 3a which clearly demonstrates that the comoving Hubble radius monotonically decreases with time and goes to zero asymptotically on both sides of the bounce. Here it may be mentioned that unlike to the comoving Hubble radius, the inverse Hubble parameter does not go to zero asymptotically but reaches to a constant value at late stage of the universe, which is depicted in 3b showing the behaviour of vs. . This corresponds to a late time accelerating universe. The asymptotic evolution of the comoving Hubble radius leads to the perturbation modes generate near the bouncing regime where the Hubble radius has an infinite size such that all the perturbation modes are contained inside the horizon. In this regard the present scenario is different from the usual matter bounce models where the Hubble radius diverges asymptotically and the perturbation modes generate far away from the bounce. Therefore in the next section we solve the perturbation equations near the bouncing point .


4.1 Scalar perturbation
The scalar metric perturbation over FRW metric can be written in the longitudinal gauge as,
(43) |
where the line element is expressed in (, ) coordinates with being the conformal time defined by and the variable symbolizes the scalar metric fluctuation. Here it may be mentioned that the spacelike and the timelike components of scalar perturbation are considered to be same, this is because the background evolution has no anisotropic stress in the present context. Moreover, we expand the radion field as,
(44) |
in terms of the background radion and the fluctuation . As a result, the scalar perturbation, in the longitudinal gauge, follows the following equations (upto first order perturbations) [127],
(45) |
where prime denotes and is the Hubble parameter in conformal time coordinate. The variation of matter energy-momentum tensor, i.e, (recall, the radion field is the only matter field in the present context) present in the right hand side of the above equations, can be obtained from 22 and are given by,
(46) |
where we explicitly used the fluctuation of the radion field shown in 44 and recall, and are the radion potential and the non-canonical kinetic term of the radion field respectively. In 46 and in the rest of the discussion the primes in and are with respect to the background radion field while the primes in and are with respect to the conformal time . Substituting the expressions of into the set of equations 45, we get
(47) |
respectively. The second of 47 can be used to obtain in terms of and , substituting which into the other two equations leads to the evolution of as,
(48) |
which explicitly depends on the non-canonical term and for , 48 reduces to that of the canonical scalar field case [127]. To solve the above perturbation equation, we will use the background evolution of the Hubble parameter and the radion field, which are obtained in the cosmic time coordinate. Thus we first transform 48 in terms of the cosmic time and for this purpose, we need the following relations,
with overdot and prime representing and respectively. As a result, 48 turns out to be,
(49) |
where is the Hubble parameter in cosmic time and recall, (with ) is the dimensionless radion field. 49 clearly reveals how the dynamics of the scalar perturbation (i.e the acceleration term ) depends on the background evolution of and . In particular, the second term in the left hand side leads to an oscillation of , the third term denotes a friction term and the fourth term indicates a restoring force. As mentioned earlier, the perturbation modes generate near the bounce and thus we are interested to solve the perturbation equations near , in which case, the background Hubble parameter and the radion field evolution follow 42. Using such background evolution of and along with the near-bounce expression of (see 32), we determine (present in the above equation) as,
(50) | |||||
where and we retain the expression of up to the leading order in . With the above expression, 49 turns out to be,
(51) |
near the bounce (i.e in the leading order of ), where and and have the following expressions,
respectively. In terms of the Fourier transformed scalar perturbation variable , 51 can be written as,
(52) |
Solving 52 for , we get
(53) |
with is the n-th order Hermite polynomial. is the integration constant which can be determined from the initial Bunch-Davies vacuum condition given by , where is the canonical Mukhanov-Sasaki variable. The Bunch-Davies vacuum choice is justified since the primordial modes at (or equivalently ) are well inside the Hubble horizon. The Bunch-Davies vacuum condition on the Mukhanov-Sasaki variable immediately leads the corresponding condition on from the following relation [127],
(54) |
and by using the background evolution of along with the expression of , we determine the initial condition of as follows,
(55) |
This makes the integration constant have the following form,
Substituting the above expression of into 53 yields the following solution for the scalar perturbation variable,
(56) |
where and are given below 51. Consequently the solution of immediately leads to the scalar power spectrum for -th modes as,
Here we would like to mention that our main aim in this section is to investigate whether the theoretical predictions of , and match with the Planck 2018 results which put a constraint on these observable quantities around the CMB scale. Therefore the scale of interest in the present context is around the CMB scale given by . With the background solution of Hubble parameter from 42, we determine the expression of the time when crosses the horizon by using the horizon crossing relation , and is given by,
(58) |
where, is the horizon crossing time of the CMB scale and recall, being the bulk curvature scale. As we will show later that the model stands to be a viable one in regard to the Planck constraints for the parameter ranges : and respectively. Such parametric ranges make the horizon crossing instance of as (the conversion may be useful). This estimation of along with 42 indicate that the scale factor, around the horizon crossing instance of , practically behaves as ; which in turn confirms the fact that the CMB scale crosses the horizon near the bouncing regime. Correspondingly, the scalar power spectrum at horizon crossing can be expressed as,
(59) |
With 59, we can determine the observable quantities like the scalar spectral index of the primordial curvature perturbations (), the scalar perturbation amplitude () etc. However before proceeding to calculate and , we will perform first the tensor perturbation, which is necessary for evaluating the tensor-to-scalar ratio ().
4.2 Tensor perturbation
In this section we consider the tensor perturbation on the FRW metric background which is defined as follows,
(60) |
where is the tensor perturbation. The variable is itself a gauge invariant quantity, and the tensor perturbed action up to quadratic order is given by [128, 129, 130],
(61) |
where , in the non-canonical scalar-tensor theory i.e the case of the present context, has the following form [128],
(62) |
61 indicates that the speed of the tensor perturbation is i.e the gravitational waves propagate with the speed of light which is unity in the natural units. This is in agreement with the event GW170817 according to which, the gravitational wave and the electromagnetic wave have the same propagation speed. At this stage, it deserves mentioning that the speed of the gravitational wave depends on the background model, as for example, the is not unity in scalar-Einstein-Gauss-Bonnet (GB) gravity theory and the deviation of from unity is proportional to the GB coupling function considered in the model. However there exists a certain class of GB coupling function for which the gravitational wave propagates with leading to the compatibility of the GB model with GW170817 (the bouncing phenomenology in such a class of Gauss-Bonnet theory which is compatible with GW170817 has been recently discussed in [69]). On other hand, the non-canonical scalar-tensor theory always leads to irrespective of the form of the non-canonical coupling function. Coming back to 62, the tensor perturbation is ensured to be stable in the present context as the condition holds. The action 61 leads to the following equation for the tensor perturbed variable ,
(63) |
The Fourier transformed tensor perturbation variable is defined as , where and represent two polarization modes. Moreover are the polarization tensors satisfying . In terms of the Fourier transformed tensor variable , 63 can be expressed as,
(64) |
The two polarization modes obey the same 64 and thus we omit the polarization index. Moreover, both the polarization modes even follow the same initial condition and hence have the same solution. Therefore, in the expression of the tensor power spectrum, we will introduce a multiplicative factor due to the contribution from both the polarization modes. As mentioned earlier, the perturbation modes generate near the bouncing regime (because at that time all the perturbation modes lie within the Hubble horizon) where the background Hubble parameter () follow the evolution presented in 42. From the solution of the form of the scale factor turns out to be which can be expanded in a Taylor series about (i.e about the bounce point) as,
We are interested to solve the perturbation near the bounce (i.e., ) where the scale factor can be approximated to be . Using this expression of the near-bounce scale factor, we determine as,
(65) |
with . Substituting this expression of into 64 and after some algebra, we get the following equation for the Fourier transformed tensor peturbation variable,
(66) |
at leading order in (since the perturbation modes generate near the bouncing phase i.e., near ). Solving 66 for , we get,
(67) |
where is an integration constant and can be determined from an initial condition. As an initial condition, we consider that the tensor perturbation field starts from the adiabatic vacuum, more precisely the initial configuration is given by, . This immediately leads to the expression of as,
(68) |
In the second equality of the above equation, we use from 62. Putting this expression of into 67 yields the final solution of as follows,
(69) |
69 represents the solution of the tensor perturbation for both the polarization modes. The solution of immediately leads to the tensor power spectrum as,
(70) | |||||
It may be noticed that and modes contribute equally to the power spectrum, as expected because their solutions behave similarly. At the horizon crossing , the tensor power spectrum turns out to be,
(71) |
with being the horizon crossing instance.
Now we can explicitly confront the model at hand with the latest Planck observational data [131], so we shall calculate the spectral index of the primordial curvature perturbations and the tensor-to-scalar ratio , which are defined as follows,
(72) |
As evident from these expressions, and are evaluated at the time of the horizon exit near the bouncing point (symbolized by ‘H.C’ in the above equations), for positive times when i.e. when the mode crosses the Hubble horizon. Using LABEL:sp16, we determine as follows,
(73) |
where is the derivative of with respect to its first argument. Therefore 73 immediately leads to the spectral index as,
(74) |
As mentioned earlier, the perturbation modes are generated and also cross the horizon near the bounce. Thus we can safely use the near-bounce scale factor in the horizon crossing condition to determine (where is the horizon crossing time). Using this relation, 74 turns out to be,
(75) |
Furthermore, the tensor-to-scalar ratio is given by,
(76) |
where the solutions of and are shown in 71 and 59 respectively. 75 and 76 clearly indicate that and depend on the dimensionless parameters and which is further connected to the Ricci scalar at horizon crossing by . Therefore, we can argue that the observable quantities and depend on and . With this information, we now directly confront the theoretical expressions of scalar spectral index 75 and tensor-to-scalar ratio 76 derived from the present model with the Planck 2018 constraints [131]. In particular, we estimate the allowed values of and which in turn can give rise to and in agreement with the Planck data. This is presented in 4 where we compute and for three choices of (viz, (blue point), (black point) and (red point) with . The allowed values of and from Planck data within and constraints are illustrated by the yellow and the blue regions respectively in 4. We note that with and all the three aforesaid values of the model estimated and are within the constraints reported by Planck 2018 data.
At this stage it may be mentioned that scalar-tensor models (with single scalar field)
which exhibit a matter bounce scenario asymptotically, such that the perturbations are generated far away
from the bouncing point deeply in the contracting regime, are
generally not consistent with the Planck results since it gives rise to an exactly scale invariant power spectrum [16].
Such inconsistency with
Planck observation was also confirmed in [68] from a slightly
different viewpoint, namely from an gravity theory. It turns out that models can be
equivalently mapped to scalar-tensor ones via conformal transformation of the metric
and, thus, the inconsistencies of the spectral index in the two different models are well justified. However there exists counter example
of this argument in the context of two field matter bounce in [24] where the authors proposed a cosmological evolution which undergoes
the phases like matter contraction, then a period of ekpyrotic contraction, followed by a non-singular bounce, and then a phase of fast roll expansion.
In such scenario, it has been showed that the primordial curvature perturbation dominated by a scale invariant component
while there are other terms which can lead to a scale dependence at small length scales, in particular there is a subdominant dependence
in the expression of scalar power spectrum. Unlike to such scenarios, here we demonstrate that a
scalar-tensor gravity model indeed leads to a viable bouncing model when the primordial perturbations are generated near the bounce.

Furthermore the scalar perturbation amplitude () is
constrained to
from the Planck results [131]. From LABEL:sp16 we note that the amplitude of scalar perturbations not only depends on and but also on the ratio of the 5D bulk curvature () and the 5D Planck mass (M) i.e . In particular, the scalar perturbation
amplitude becomes when we take and . This is in accordance with the Planck constraints mentioned above provided lies within such that the bulk curvature is constrained to be less than the 5D Planck mass, which in turn confirms the validity of the background classical solution.
However, it may be mentioned that the allowed range of is sensitive to the choice of , i.e. a different will lead to a
different allowed range for the parameter . As an example, leads to the scalar perturbation
amplitude as which becomes consistent with the Planck results
for . However with the condition , the assumption of the background classical solution ceases to hold true, which is not desirable. Therefore, as a whole, the
observable quantities , and are simultaneously
compatible with the Planck constraints for the parameter ranges :
, , respectively. Such parametric ranges make the
horizon crossing Ricci scalar of the order .
Before concluding, we would like to mention that the present paper studies a non-singular bounce from a warped braneworld scenario with dynamical branes, which is found to yield a nearly scale-invariant power spectra of primordial perturbations. However, in the background of the contracting era, the anisotropy grows with the scale factor as and thus the contracting stage becomes unstable to the growth of anisotropies, which is known as the BKL instability [132]. Thus similar to many other bounce models, except the ekpyrotic bounce scenario [24, 26, 133, 134], the present model also suffers from the BKL instability. Thereby it may be an interesting study to explore the possible effects of radion dynamics in an ekpyrotic bounce scenario to avoid the BKL instability. This however may be considered in a future work.
5 Conclusion
We consider a five dimensional warped braneworld scenario with two 3-branes embedded within the 5D spacetime, where the branes have a non-zero cosmological constant leading to a non-flat brane geometry. With dynamical branes the interbrane distance is treated as a scalar field, the so-called radion or the modulus, which generates its own potential when the effective action is obtained as a consequence of compactification of the extra coordinate. Such a radion field is also associated with a non-canonical kinetic term at the level of the four dimesional effective action which exhibits a transition from a normal to a phantom regime (i.e from to ) as the radion field goes from higher to lower values. With the vanishing of the brane cosmological constant , the branes become flat such that the radion potential ceases to exist while the radion kinetic term becomes canonical. Such a non-flat warped braneworld scenario is important as it can simultaneously address the gauge-hierarchy problem and the stabilization of the modulus without the necessity of any additional scalar field of unknown origin.
The presence of the phantom regime is further interesting as the cosmological evolution of the radion field in the FRW background leads to a violation of the null energy condition, necessary to ensure a non-singular bounce in our visible universe. This motivates us to explore the prospect of the radion field in triggering a bouncing universe which in turn can potentially avoid the Big-Bang singularity. Note that the radion field by which the bounce is driven arises naturally from compactification in the effective four-dimensional theory and generates its own potential due to the presence of the brane cosmological constant, unlike most of the scalar-tensor bounce models where the scalar potentials are constructed by hand to explain the observations and often their origin remains unexplained.
An analysis of the background cosmological evolution of the Hubble parameter and the radion field reveals that the radion field starts its journey from the normal regime (i.e regime) and decreases monotonically in magnitude with cosmic time until it transits to the phantom era where the bounce occurs. With further time evolution the radion asymptotically stabilizes to the value which also represents the inflection point of the modulus potential. Such an asymptotic magnitude of the radion field can stabilize the modulus to the appropriate value where the gauge-hierarchy issue can also be adequately addressed.
With the background evolution, we further investigate the cosmological evolution of the scalar and tensor perturbations to the FRW metric from the present model. The primordial perturbation modes in the present context generate near the bounce because at that time the relevant perturbation modes are within the horizon, unlike the usual matter bounce scenario where the perturbation modes generate deeply in the contracting regime far away from the bouncing point. As a result the tensor perturbation is found to be suppressed in comparison to the scalar perturbation and the ratio of tensor to scalar perturbation amplitude becomes less than unity in accordance with the Planck results. Moreover, the speed of propagation of the tensor perturbation turns out to be the same as the speed of light, in agreement with the event GW170817. We compute the scalar spectral index , the tensor to scalar ratio and the amplitude of the scalar perturbations from the present model which turns out to be pleasantly in agreement with the latest Planck 2018 observations, well within the 1- regime.
Acknowledgments
This research was partially supported in part by the International Centre for Theoretical Sciences (ICTS) for the program - Physics of the Early Universe - An Online Precursor (code: ICTS/peu2020/08).
References
- [1] A.H. Guth; Phys.Rev. D23 347-356 (1981).
- [2] A. D. Linde, Contemp. Concepts Phys. 5 (1990) 1 [hep-th/0503203].
- [3] D. Langlois, hep-th/0405053.
- [4] A. Riotto, ICTP Lect. Notes Ser. 14 (2003) 317 [hep-ph/0210162].
- [5] J. D. Barrow and P. Saich; Class. Quantum Grav. 10, 279 (1993).
- [6] J. D. Barrow and J. P. Mimoso; Phys. Rev. D 50, 3746 (1994).
- [7] N. Banerjee, S. Sen; Phys.Rev. D57, 4614 (1998).
- [8] D. Baumann, doi:10.1142/9789814327183 0010 [arXiv:0907.5424 [hep-th]].
- [9] R. H. Brandenberger, arXiv:1206.4196 [astro-ph.CO].
- [10] R. Brandenberger and P. Peter, arXiv:1603.05834 [hep-th].
- [11] D. Battefeld and P. Peter, Phys. Rept. 571 (2015) 1 doi:10.1016/j.physrep.2014.12.004 [arXiv:1406.2790 [astro-ph.CO]].
- [12] M. Novello and S. E. P. Bergliaffa, “Bouncing Cosmologies,” Phys. Rept. 463 (2008) 127 doi:10.1016/j.physrep.2008.04.006 [arXiv:0802.1634 [astro-ph]].
- [13] Y. F. Cai, Sci. China Phys. Mech. Astron. 57 (2014) 1414 doi:10.1007/s11433-014-5512-3 [arXiv:1405.1369 [hep-th]].
- [14] Y. Cai, Y. Wan, H. G. Li, T. Qiu and Y. S. Piao, JHEP 01 (2017), 090 doi:10.1007/JHEP01(2017)090 [arXiv:1610.03400 [gr-qc]].
- [15] Y. Cai, H. G. Li, T. Qiu and Y. S. Piao, Eur. Phys. J. C 77 (2017) no.6, 369 doi:10.1140/epjc/s10052-017-4938-y [arXiv:1701.04330 [gr-qc]].
- [16] J. de Haro and Y. F. Cai, Gen. Rel. Grav. 47 (2015) no.8, 95 doi:10.1007/s10714-015-1936-y [arXiv:1502.03230 [gr-qc]].
- [17] J. L. Lehners, Class. Quant. Grav. 28 (2011) 204004 doi:10.1088/0264-9381/28/20/204004 [arXiv:1106.0172 [hep-th]].
- [18] J. L. Lehners, Phys. Rept. 465 (2008) 223 doi:10.1016/j.physrep.2008.06.001 [arXiv:0806.1245 [astro-ph]].
- [19] Y. K. E. Cheung, C. Li and J. D. Vergados, arXiv:1611.04027 [astro-ph.CO].
- [20] Y. F. Cai, A. Marciano, D. G. Wang and E. Wilson-Ewing, Universe 3 (2016) no.1, 1 doi:10.3390/Universe3010001 [arXiv:1610.00938 [astro-ph.CO]].
- [21] C. Cattoen and M. Visser, Class. Quant. Grav. 22 (2005) 4913 doi:10.1088/0264-9381/22/23/001 [gr-qc/0508045].
- [22] C. Li, R. H. Brandenberger and Y. K. E. Cheung, Phys. Rev. D 90 (2014) no.12, 123535 doi:10.1103/PhysRevD.90.123535 [arXiv:1403.5625 [gr-qc]].
- [23] D. Brizuela, G. A. D. Mena Marugan and T. Pawlowski, Class. Quant. Grav. 27 (2010) 052001 doi:10.1088/0264-9381/27/5/052001 [arXiv:0902.0697 [gr-qc]].
- [24] Y. F. Cai, E. McDonough, F. Duplessis and R. H. Brandenberger, JCAP 1310 (2013) 024 doi:10.1088/1475-7516/2013/10/024 [arXiv:1305.5259 [hep-th]].
- [25] J. Quintin, Y. F. Cai and R. H. Brandenberger, Phys. Rev. D 90 (2014) no.6, 063507 doi:10.1103/PhysRevD.90.063507 [arXiv:1406.6049 [gr-qc]].
- [26] Y. F. Cai, R. Brandenberger and P. Peter, Class. Quant. Grav. 30 (2013) 075019 doi:10.1088/0264-9381/30/7/075019 [arXiv:1301.4703 [gr-qc]].
- [27] N. J. Poplawski, Phys. Rev. D 85 (2012) 107502 doi:10.1103/PhysRevD.85.107502 [arXiv:1111.4595 [gr-qc]].
- [28] M. Koehn, J. L. Lehners and B. Ovrut, Phys. Rev. D 93 (2016) no.10, 103501 doi:10.1103/PhysRevD.93.103501 [arXiv:1512.03807 [hep-th]].
- [29] S. D. Odintsov and V. K. Oikonomou, Phys. Rev. D 92 (2015) no.2, 024016 doi:10.1103/PhysRevD.92.024016 [arXiv:1504.06866 [gr-qc]].
- [30] S. Nojiri, S. D. Odintsov and V. K. Oikonomou, Phys. Rev. D 93 (2016) no.8, 084050 doi:10.1103/PhysRevD.93.084050 [arXiv:1601.04112 [gr-qc]].
- [31] S. D. Odintsov and V. K. Oikonomou, arXiv:1512.04787 [gr-qc].
- [32] S. D. Odintsov, V. K. Oikonomou and T. Paul, Class. Quant. Grav. 37 (2020) no.23, 235005 doi:10.1088/1361-6382/abbc47 [arXiv:2009.09947 [gr-qc]].
- [33] M. Koehn, J. L. Lehners and B. A. Ovrut, Phys. Rev. D 90 (2014) no.2, 025005 doi:10.1103/PhysRevD.90.025005 [arXiv:1310.7577 [hep-th]].
- [34] L. Battarra and J. L. Lehners, JCAP 1412 (2014) no.12, 023 doi:10.1088/1475-7516/2014/12/023 [arXiv:1407.4814 [hep-th]].
- [35] J. Martin, P. Peter, N. Pinto Neto and D. J. Schwarz, Phys. Rev. D 65 (2002) 123513 doi:10.1103/PhysRevD.65.123513 [hep-th/0112128].
- [36] J. Khoury, B. A. Ovrut, P. J. Steinhardt and N. Turok, Phys. Rev. D 64 (2001) 123522 doi:10.1103/PhysRevD.64.123522 [hep-th/0103239].
- [37] E. I. Buchbinder, J. Khoury and B. A. Ovrut, Phys. Rev. D 76 (2007) 123503 doi:10.1103/PhysRevD.76.123503 [hep-th/0702154].
- [38] M. G. Brown, K. Freese and W. H. Kinney, JCAP 0803 (2008) 002 doi:10.1088/1475-7516/2008/03/002 [astro-ph/0405353].
- [39] J. C. Hackworth and E. J. Weinberg, Phys. Rev. D 71 (2005) 044014 doi:10.1103/PhysRevD.71.044014 [hep-th/0410142].
- [40] S. Nojiri and S. D. Odintsov, Phys. Lett. B 637 (2006) 139 doi:10.1016/j.physletb.2006.04.026 [hep-th/0603062].
- [41] M. C. Johnson and J. L. Lehners, Phys. Rev. D 85 (2012) 103509 doi:10.1103/PhysRevD.85.103509 [arXiv:1112.3360 [hep-th]].
- [42] P. Peter and N. Pinto-Neto, Phys. Rev. D 66 (2002) 063509 doi:10.1103/PhysRevD.66.063509 [hep-th/0203013].
- [43] M. Gasperini, M. Giovannini and G. Veneziano, Phys. Lett. B 569 (2003) 113 doi:10.1016/j.physletb.2003.07.028 [hep-th/0306113].
- [44] P. Creminelli, A. Nicolis and M. Zaldarriaga, Phys. Rev. D 71 (2005) 063505 doi:10.1103/PhysRevD.71.063505 [hep-th/0411270].
- [45] J. L. Lehners and E. Wilson-Ewing, JCAP 1510 (2015) no.10, 038 doi:10.1088/1475-7516/2015/10/038 [arXiv:1507.08112 [astro-ph.CO]].
- [46] J. Mielczarek, M. Kamionka, A. Kurek and M. Szydlowski, JCAP 1007 (2010) 004 doi:10.1088/1475-7516/2010/07/004 [arXiv:1005.0814 [gr-qc]].
- [47] J. L. Lehners and P. J. Steinhardt, Phys. Rev. D 87 (2013) no.12, 123533 doi:10.1103/PhysRevD.87.123533 [arXiv:1304.3122 [astro-ph.CO]].
- [48] Y. F. Cai, J. Quintin, E. N. Saridakis and E. Wilson-Ewing, JCAP 1407 (2014) 033 doi:10.1088/1475-7516/2014/07/033 [arXiv:1404.4364 [astro-ph.CO]].
- [49] Y. F. Cai, T. Qiu, Y. S. Piao, M. Li and X. Zhang, JHEP 0710 (2007) 071 doi:10.1088/1126-6708/2007/10/071 [arXiv:0704.1090 [gr-qc]].
- [50] Y. F. Cai and E. N. Saridakis, Class. Quant. Grav. 28 (2011) 035010 doi:10.1088/0264-9381/28/3/035010 [arXiv:1007.3204 [astro-ph.CO]].
- [51] P. P. Avelino and R. Z. Ferreira, Phys. Rev. D 86 (2012) 041501 doi:10.1103/PhysRevD.86.041501 [arXiv:1205.6676 [astro-ph.CO]].
- [52] J. D. Barrow, D. Kimberly and J. Magueijo, Class. Quant. Grav. 21 (2004) 4289 doi:10.1088/0264-9381/21/18/001 [astro-ph/0406369].
- [53] J. Haro and E. Elizalde, JCAP 1510 (2015) no.10, 028 doi:10.1088/1475-7516/2015/10/028 [arXiv:1505.07948 [gr-qc]].
- [54] E. Elizalde, J. Haro and S. D. Odintsov, Phys. Rev. D 91 (2015) no.6, 063522 doi:10.1103/PhysRevD.91.063522 [arXiv:1411.3475 [gr-qc]].
- [55] A. Das, D. Maity, T. Paul and S. SenGupta, Eur. Phys. J. C 77 (2017) no.12, 813 doi:10.1140/epjc/s10052-017-5396-2 [arXiv:1706.00950 [hep-th]].
- [56] J. de Haro, JCAP 1211 (2012) 037 [arXiv:1207.3621 [gr-qc]].
- [57] E. Wilson-Ewing, JCAP 1303 (2013) 026 doi:10.1088/1475-7516/2013/03/026 [arXiv:1211.6269 [gr-qc]].
- [58] Y. F. Cai, T. t. Qiu, R. Brandenberger and X. m. Zhang, Phys. Rev. D 80 (2009) 023511 doi:10.1103/PhysRevD.80.023511 [arXiv:0810.4677 [hep-th]].
- [59] F. Finelli and R. Brandenberger, Phys. Rev. D 65 (2002) 103522 doi:10.1103/PhysRevD.65.103522 [hep-th/0112249].
- [60] Y. F. Cai, R. Brandenberger and X. Zhang, Phys. Lett. B 703 (2011) 25 doi:10.1016/j.physletb.2011.07.074 [arXiv:1105.4286 [hep-th]].
- [61] J. Haro and J. Amorós, PoS FFP 14 (2016) 163 doi:10.22323/1.224.0163 [arXiv:1501.06270 [gr-qc]].
- [62] Y. F. Cai, R. Brandenberger and X. Zhang, JCAP 1103 (2011) 003 doi:10.1088/1475-7516/2011/03/003 [arXiv:1101.0822 [hep-th]].
- [63] J. Haro and J. Amoros, JCAP 1412 (2014) no.12, 031 doi:10.1088/1475-7516/2014/12/031 [arXiv:1406.0369 [gr-qc]].
- [64] R. Brandenberger, Phys. Rev. D 80 (2009) 043516 doi:10.1103/PhysRevD.80.043516 [arXiv:0904.2835 [hep-th]].
- [65] J. de Haro and J. Amoros, JCAP 1408 (2014) 025 doi:10.1088/1475-7516/2014/08/025 [arXiv:1403.6396 [gr-qc]].
- [66] S. D. Odintsov and V. K. Oikonomou, Phys. Rev. D 90 (2014) no.12, 124083 doi:10.1103/PhysRevD.90.124083 [arXiv:1410.8183 [gr-qc]].
- [67] T. Qiu and K. C. Yang, JCAP 1011 (2010) 012 doi:10.1088/1475-7516/2010/11/012 [arXiv:1007.2571 [astro-ph.CO]].
- [68] S. Nojiri, S. D. Odintsov, V. K. Oikonomou and T. Paul, Phys. Rev. D 100 (2019) no.8, 084056 doi:10.1103/PhysRevD.100.084056 [arXiv:1910.03546 [gr-qc]].
- [69] E. Elizalde, S. D. Odintsov, V. K. Oikonomou and T. Paul, Nucl. Phys. B 954 (2020), 114984 doi:10.1016/j.nuclphysb.2020.114984 [arXiv:2003.04264 [gr-qc]].
- [70] K. Bamba, J. de Haro and S. D. Odintsov, JCAP 1302 (2013) 008 doi:10.1088/1475-7516/2013/02/008 [arXiv:1211.2968 [gr-qc]].
- [71] G. F. R. Ellis, J. Murugan and C. G. Tsagas, Class. Quant. Grav. 21 (2004) no.1, 233-250 doi:10.1088/0264-9381/21/1/016 [arXiv:gr-qc/0307112 [gr-qc]].
- [72] B. C. Paul, S. D. Maharaj and A. Beesham, [arXiv:2008.00169 [astro-ph.CO]].
- [73] S. L. Li, H. Lü, H. Wei, P. Wu and H. Yu, Phys. Rev. D 99 (2019) no.10, 104057 doi:10.1103/PhysRevD.99.104057 [arXiv:1903.03940 [gr-qc]].
- [74] Y. F. Cai, M. Li and X. Zhang, Phys. Lett. B 718 (2012), 248-254 doi:10.1016/j.physletb.2012.10.065 [arXiv:1209.3437 [hep-th]].
- [75] S. Dutta, S. Mukerjiand and S. Chakraborty, Adv. High Energy Phys. 2016 (2016), 7404218 doi:10.1155/2016/7404218
- [76] S. Bag, V. Sahni, Y. Shtanov and S. Unnikrishnan, JCAP 07 (2014), 034 doi:10.1088/1475-7516/2014/07/034 [arXiv:1403.4243 [astro-ph.CO]].
- [77] R. H. Brandenberger, [arXiv:0808.0746 [hep-th]].
- [78] R. H. Brandenberger, Class. Quant. Grav. 28 (2011), 204005 doi:10.1088/0264-9381/28/20/204005 [arXiv:1105.3247 [hep-th]].
- [79] R. H. Brandenberger and C. Vafa, Nucl. Phys. B 316 (1989), 391-410 doi:10.1016/0550-3213(89)90037-0
- [80] J. Kripfganz and H. Perlt, Class. Quant. Grav. 5 (1988), 453 doi:10.1088/0264-9381/5/3/006
- [81] T. Battefeld and S. Watson, Rev. Mod. Phys. 78 (2006), 435-454 doi:10.1103/RevModPhys.78.435 [arXiv:hep-th/0510022 [hep-th]].
- [82] A. Nayeri, R. H. Brandenberger and C. Vafa, Phys. Rev. Lett. 97 (2006), 021302 doi:10.1103/PhysRevLett.97.021302 [arXiv:hep-th/0511140 [hep-th]].
- [83] R. H. Brandenberger, A. Nayeri, S. P. Patil and C. Vafa, Int. J. Mod. Phys. A 22 (2007), 3621-3642 doi:10.1142/S0217751X07037159 [arXiv:hep-th/0608121 [hep-th]].
- [84] J. Polchinski, String theory. Vol. 1: An introduction to the bosonic string, 1998
- [85] J. Polchinski, String theory. Vol. 2: Superstring theory and beyon, 1998
- [86] P. Horava and E. Witten, Nucl. Phys. B 460 (1996), 506-524 doi:10.1016/0550-3213(95)00621-4 [arXiv:hep-th/9510209 [hep-th]].
- [87] P. Horava and E. Witten, Nucl. Phys. B 475 (1996), 94-114 doi:10.1016/0550-3213(96)00308-2 [arXiv:hep-th/9603142 [hep-th]].
- [88] I. Antoniadis, Phys. Lett. B 246 (1990), 377-384 doi:10.1016/0370-2693(90)90617-F
- [89] J. D. Lykken, Phys. Rev. D 54 (1996), 3693-3697 doi:10.1103/PhysRevD.54.R3693 [arXiv:hep-th/9603133 [hep-th]].
- [90] Z. Kakushadze and S. H. H. Tye, Nucl. Phys. B 548 (1999), 180-204 doi:10.1016/S0550-3213(99)00082-6 [arXiv:hep-th/9809147 [hep-th]].
- [91] N. Arkani-Hamed, S. Dimopoulos and G. R. Dvali, Phys. Lett. B 429 (1998), 263-272 doi:10.1016/S0370-2693(98)00466-3 [arXiv:hep-ph/9803315 [hep-ph]].
- [92] I. Antoniadis, N. Arkani-Hamed, S. Dimopoulos and G. R. Dvali, Phys. Lett. B 436 (1998), 257-263 doi:10.1016/S0370-2693(98)00860-0 [arXiv:hep-ph/9804398 [hep-ph]].
- [93] N. Arkani-Hamed, S. Dimopoulos and G. R. Dvali, Phys. Rev. D 59 (1999), 086004 doi:10.1103/PhysRevD.59.086004 [arXiv:hep-ph/9807344 [hep-ph]].
- [94] L. Randall and R. Sundrum, Phys. Rev. Lett. 83 (1999), 3370-3373 doi:10.1103/PhysRevLett.83.3370 [arXiv:hep-ph/9905221 [hep-ph]].
- [95] L. Randall and R. Sundrum, Phys. Rev. Lett. 83 (1999), 4690-4693 doi:10.1103/PhysRevLett.83.4690 [arXiv:hep-th/9906064 [hep-th]].
- [96] J. D. Lykken and L. Randall, JHEP 06 (2000), 014 doi:10.1088/1126-6708/2000/06/014 [arXiv:hep-th/9908076 [hep-th]].
- [97] N. Kaloper, Phys. Rev. D 60 (1999), 123506 doi:10.1103/PhysRevD.60.123506 [arXiv:hep-th/9905210 [hep-th]].
- [98] S. Das, D. Maity and S. SenGupta, JHEP 05 (2008), 042 doi:10.1088/1126-6708/2008/05/042 [arXiv:0711.1744 [hep-th]].
- [99] Csaki, C.; Graesser, M.L.; Kribs, G.D. Phys. Rev. D 63 (2001), 065002. doi:10.1103/PhysRevD.63.065002.
- [100] DeWolfe, O.; Freedman, D.Z.; Gubser, S.S.; Karch, A. Phys. Rev. D 62 (2000), 046008. doi:10.1103/PhysRevD.62.046008.
- [101] Lesgourgues, J.; Pastor, S.; Peloso, M.; Sorbo, L. Phys. Lett. B 489 (2000), 411. doi:10.1016/S0370-2693(00)00943-6.
- [102] Csaki, C.; Graesser, M.; Randall, L.; Terning, J. Phys. Rev. D 62 (2000), 045015. doi:10.1103/PhysRevD.62.045015.
- [103] Binetruy, P.; Deffayet, C.; Langlois, D. Nucl. Phys. B 2000 565 (2000), 269–287. doi:10.1016/S0550-3213(99)00696-3.
- [104] Csaki, C.; Graesser, M.; Kolda, C.F.; Terning, J. Phys. Lett. B 462 (1999), 34–40. doi:10.1016/S0370-2693(99)00896-5.
- [105] Cline, J.M. Cosmological expansion in the Randall-Sundrum warped compactification. arXiv 2009, arXiv:0001285.
- [106] Nojiri, S.; Odintsov, S.D. JHEP 7 (2000), 49. doi:10.1088/1126-6708/2000/07/049.
- [107] Nojiri, S.; Odintsov, S.D.; Ogushi, S. Phys. Rev. D 65 (2002), 023521. doi:10.1103/PhysRevD.65.023521.
- [108] Davoudiasl, H.; Hewett, J.L.; Rizzo, T.G. Phys. Rev. Lett. 84 (2000), 2080. doi:10.1103/PhysRevLett.84.2080.
- [109] Das, A.; SenGupta, S. Eur. Phys. J. C 76 (2016), 423. doi:10.1140/epjc/s10052-016-4264-9.
- [110] Tang, Y. JHEP 8 (2012), 78. doi:10.1007/JHEP08(2012)078.
- [111] Arun, M.T.; Choudhury, D.; Das, A.; SenGupta, S. Phys. Lett. B 746 (2015), 266–275. doi:10.1016/j.physletb.2015.05.008.
- [112] Das, A.; SenGupta, S. 126 GeV Higgs and ATLAS bound on the lightest graviton mass in Randall-Sundrum model. arXiv 2013, arXiv:1303.2512.
- [113] Banerjee, I.; Chakraborty, S.; SenGupta, S. Phys. Rev. D 99 (2019), 023515. doi:10.1103/PhysRevD.99.023515.
- [114] Chakraborty, S.; Sengupta, S. Eur. Phys. J. C 74 (2014), 3045. doi:10.1140/epjc/s10052-014-3045-6.
- [115] T. Paul and S. Sengupta, Phys. Rev. D 93 (2016) no.8, 085035 doi:10.1103/PhysRevD.93.085035 [arXiv:1601.05564 [hep-ph]].
- [116] Banerjee, N.; Paul, T. Eur. Phys. J. C 77 (2017), 672. doi:10.1140/epjc/s10052-017-5256-0.
- [117] J. Mitra, T. Paul and S. SenGupta, Eur. Phys. J. C 77 (2017) no.12, 833 doi:10.1140/epjc/s10052-017-5420-6 [arXiv:1707.06532 [hep-th]].
- [118] E. Elizalde, S. D. Odintsov, T. Paul and D. Sáez-Chillón Gómez, Phys. Rev. D 99 (2019) no.6, 063506 doi:10.1103/PhysRevD.99.063506 [arXiv:1811.02960 [gr-qc]].
- [119] K. Aditya, I. Banerjee, A. Banerjee and S. SenGupta, doi:10.1093/mnras/staa3104 [arXiv:2004.05627 [gr-qc]].
- [120] I. Banerjee, S. Chakraborty and S. SenGupta, Phys. Rev. D 101 (2020) no.4, 041301 doi:10.1103/PhysRevD.101.041301 [arXiv:1909.09385 [gr-qc]].
- [121] I. Banerjee, S. Chakraborty and S. SenGupta, Phys. Rev. D 100 (2019) no.4, 044045 doi:10.1103/PhysRevD.100.044045 [arXiv:1905.08043 [gr-qc]].
- [122] N. Aghanim et al. [Planck], Astron. Astrophys. 641 (2020), A6 doi:10.1051/0004-6361/201833910 [arXiv:1807.06209 [astro-ph.CO]].
- [123] W. D. Goldberger and M. B. Wise, Phys. Rev. Lett. 83 (1999), 4922-4925 doi:10.1103/PhysRevLett.83.4922 [arXiv:hep-ph/9907447 [hep-ph]].
- [124] W. D. Goldberger and M. B. Wise, Phys. Lett. B 475 (2000), 275-279 doi:10.1016/S0370-2693(00)00099-X [arXiv:hep-ph/9911457 [hep-ph]].
- [125] I. Banerjee and S. SenGupta, Eur. Phys. J. C 77 (2017) no.5, 277 doi:10.1140/epjc/s10052-017-4857-y [arXiv:1705.05015 [hep-th]].
- [126] M. Sasaki, T. Shiromizu and K. i. Maeda, Phys. Rev. D 62 (2000), 024008 doi:10.1103/PhysRevD.62.024008 [arXiv:hep-th/9912233 [hep-th]].
- [127] R. H. Brandenberger, Lect. Notes Phys. 646 (2004), 127-167 doi:10.1007/978-3-540-40918-2_5 [arXiv:hep-th/0306071 [hep-th]].
- [128] J. c. Hwang and H. Noh, Phys. Rev. D 71 (2005) 063536 doi:10.1103/PhysRevD.71.063536 [gr-qc/0412126].
- [129] H. Noh and J. c. Hwang, Phys. Lett. B 515 (2001) 231 doi:10.1016/S0370-2693(01)00875-9 [astro-ph/0107069].
- [130] J. c. Hwang and H. Noh, Phys. Rev. D 66 (2002) 084009 doi:10.1103/PhysRevD.66.084009 [hep-th/0206100].
- [131] Y. Akrami et al. [Planck Collaboration], arXiv:1807.06211 [astro-ph.CO].
- [132] V.A. Belinskii, I.M. Khalatnikov and E.M. Lifshitz ; Advances in Physics 19, 525 (1970).
- [133] J. K. Erickson, D. H. Wesley, P. J. Steinhardt and N. Turok, Phys. Rev. D 69 (2004) 063514 doi:10.1103/PhysRevD.69.063514 [hep-th/0312009].
- [134] D. Garfinkle, W. C. Lim, F. Pretorius and P. J. Steinhardt, Phys. Rev. D 78 (2008) 083537 doi:10.1103/PhysRevD.78.083537 [arXiv:0808.0542 [hep-th]].