This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Cosmological perturbation theory using generalized Einstein de Sitter cosmologies

Michael Joyce joyce@lpnhe.in2p3.fr Laboratoire de Physique Nucléaire et de Hautes Energies, UPMC IN2P3 CNRS UMR 7585, Sorbonne Université, 4, place Jussieu, 75252 Paris Cedex 05, France    Azrul Pohan azrul.pohan@lpnhe.in2p3.fr Laboratoire de Physique Nucléaire et de Hautes Energies, UPMC IN2P3 CNRS UMR 7585, Sorbonne Université, 4, place Jussieu, 75252 Paris Cedex 05, France Laboratory of Theoretical Physics, Institut Teknologi Sumatera, Lampung 35365, Indonesia
Abstract

The separable analytical solution in standard perturbation theory for an Einstein de Sitter (EdS) universe can be generalized to the wider class of such cosmologies (“generalized EdS”, or gEdS) in which a fraction of the pressure-less fluid does not cluster. We derive the corresponding kernels in both Eulerian perturbation theory (EPT) and Lagrangian perturbation theory, generalizing the canonical EdS expressions to a one-parameter family where the parameter can be taken to be the exponent α\alpha of the growing mode linear amplification D(a)aαD(a)\propto a^{\alpha}. For the power spectrum (PS) at one loop in EPT, the contribution additional to standard EdS is given, for each of the ‘13’ and ‘22’ terms, as a function of two infra-red safe integrals. In the second part of the paper we show that the calculation of cosmology-dependent corrections in perturbation theory in standard (e.g. LCDM-like) models can be simplified, and their magnitude and parameter dependence better understood, by relating them to our analytic results for gEdS models. At second order the time dependent kernels are equivalent to the analytic kernels of the gEdS model with α\alpha replaced by a single redshift dependent effective growth rate α2(z)\alpha_{2}(z). At third order the time evolution can be conveniently parametrized in terms of two additional such effective growth rates. For the PS calculated at one loop order, the correction to the PS relative to the EdS limit can be expressed in terms of just α2(z)\alpha_{2}(z), one additional effective growth rate function and the four infra-red safe integrals of the gEdS limit. This is much simplified compared to expressions in the literature that use six or eight red-shift dependent functions and are not explicitly infra-red safe. Using the analytic gEdS expression for the PS with α=α2(z)\alpha=\alpha_{2}(z) gives a good approximation (to 25%\sim 25\%) for the exact result.

preprint: APS/123-QED

I Introduction

Understanding the origin of the large-scale structures in the Universe first revealed several decades ago by early red-shift surveys remains a central and open problem in cosmology. This is driven by the ever-increasing richness and accuracy of observational data which will grow in the coming years with programs such as Euclid, DESI and LSST [1, 2, 3]. In standard cosmological models the clustering observed at cosmological scales today is the result of evolution under the gravity of primordial fluctuations, that is highly constrained by observations of the fluctuations in the cosmic microwave background, with potential fluctuations 105\sim 10^{-5} at the time of decoupling. Calculation of predictions derived from the non-linear equations governing the evolution of matter fluctuations rests in practice essentially on numerical solutions using the NN-body method (for a recent review, see e.g. [4]). At sufficiently early times and/or large scales, when fluctuations to uniformity are small, analytical perturbative approaches may be used and provide an invaluable guideline and control on numerical simulations. Perturbation theory beyond the leading (linear) order was originally treated in early works (e.g. [5, 6, 7, 8, 9, 10, 11, 12, 13, 14]) and has been developed since in an extensive literature (for a review, see [14]). In recent years perturbation theory has remained highly topical, with development of various different aspects and cosmological applications (see e.g. [15, 16, 17, 18, 19, 20, 21, 22]), and, more recently, of the so-called Effective Field Theory (EFT) of cosmological structure formation (see e.g. [15, 23, 24, 25, 26]).

In this article we explore and derive results for the perturbation theory of cosmological structure formation in a particular set of idealized cosmological models. We refer to these as “generalized Einstein de Sitter” (gEdS) models as they are models in which the background cosmological evolution is that of an EdS cosmology. However, only a fraction of the total energy density driving the expansion of the background corresponds to clustering matter, the rest remaining uniform. This is simply equivalent to a smooth matter-like component (e.g. arising from a homogeneous mode of a scalar field, or from a massive neutrino-like contribution when perturbations are neglected). We derive analytical solutions for the key quantities (kernels) in both standard (Eulerian) perturbation theory (SPT) and also in Lagrangian perturbation theory (LPT), and use them to derive the expression for the power spectrum at the leading non-trivial order (one loop). The motivation for studying these models is twofold. On the one hand, as we will show in the second part of this paper, these models turn out to be very useful to better understand, and simplify the calculation of, the small cosmology-dependent corrections to results obtained using the usual “separable EdS” approximation made in applying perturbation theory to the standard (i.e. LCDM-like) cosmologies [27, 28, 29, 30, 20, 31, 32, 14]. In particular we will show how expressions for the non-EdS corrections to the one-loop PS in perturbation theory can be given (exactly) in terms of just two redshift dependent “effective growth rates”, and, to a good approximation, in terms of a single such function. On the other hand, our motivation comes from the intrinsic interest of new analytical results in perturbation theory in order to test its validity, and in particular to probe its accuracy in predicting the dependence on expansion history beyond the usually used approximation. The models we focus on are particularly suitable for robust numerical tests as in the case of initial Gaussian fluctuations specified by a simple power-law power spectrum these gEdS models are scale-free and characterized (at least for some range of spectra) by “self-similarity” in their evolution [33, 34]. In principle this property allows one to obtain very accurate numerical results with which to test perturbation theory (in particular by applying the methods recently discussed for scale-free EdS models in [35, 36]). An analysis of these “generalized scale-free” models and numerical tests will be reported in a separate forthcoming article.

The article is organized as follows. The next section presents first the family of gEdS models, and then the calculation of perturbation theory kernels in them for both Eulerian and Lagrangian formulations, giving the density and velocity kernels up to the third order. In section 3 we study the convergence properties of the one-loop PS in gEdS cosmologies, considering how the well-known results for standard EdS are modified. Section 4 considers how the functions characterising the exact evolution of the PS in perturbation theory in a class of standard (LCDM-like) cosmological models can be approximated by interpolation of gEdS models. In the final section we summarize and comment on further possible exploitation of the results we have derived, notably in relation to understanding better the ultra-violet divergences in standard perturbation theory. Some details of our calculations are relegated to appendices, as well as additional detailed comparison with the previous results of [27], and also to those of [11, 20, 32]. For readers wishing to make use of our simplified expression for the cosmological corrections to the EdS approximation for the one-loop PS in standard cosmologies, given in terms of just two redshift-dependent functions, without going through the full derivation given in the body of the article, we give the necessary equations also in a short appendix.

II Perturbation Theory kernels in generalized EdS models

II.1 Generalized EdS models

We consider perturbed Friedman-Lemaitre-Robertson-Walker (FLRW) models with a pressureless perturbed matter component and a smooth component (i.e. without perturbations). The Friedmann equations can be written as

2=8πG3a2[ρm+ρs],τ=4πG3a2[ρm+(1+3ws)ρs],\begin{split}\mathcal{H}^{2}&=\frac{8\pi G}{3}a^{2}\big{[}\rho_{m}+\rho_{s}\big{]},\\ \frac{\partial{\mathcal{H}}}{\partial{\tau}}&=-\frac{4\pi G}{3}a^{2}\big{[}\rho_{m}+(1+3w_{s})\rho_{s}\big{]},\end{split} (1)

where =d(loga)/dτ\mathcal{H}=d(\log a)/d\tau and τ\tau is the conformal time related to cosmic time tt by τ=𝑑t/a\tau=\int dt/a, aa is the scale factor, ρm\rho_{m} is the mean density of the perturbed matter component, ρs\rho_{s} is the smooth “dark energy” component with equation of state ps=wsρsp_{s}=w_{s}\rho_{s} (where wsw_{s} can be a function of aa). The case ws=1w_{s}=-1 corresponds to the Lambda Cold Dark Matter (LCDM) cosmology. As canonically, we define the fraction of matter and dark energy as

Ωm=ρmρtot,Ωs=ρsρtot,\Omega_{m}=\frac{\rho_{m}}{\rho_{tot}},\quad\Omega_{s}=\frac{\rho_{s}}{\rho_{tot}}, (2)

where ρtot=ρm+ρs\rho_{tot}=\rho_{m}+\rho_{s} is the total energy density (and Ωm+Ωs=1\Omega_{m}+\Omega_{s}=1). Once wsw_{s} is specified, the model is fully characterized by specifying the value at some time of one of these two parameters. Our analysis of standard cosmologies below applies to this class of models, with the sole further assumption that the total energy density asymptotically scales as matter at early times.

The family of cosmologies we will focus on first, those we will refer to as generalized Einstein de Sitter cosmologies (hereafter gEdS), simply correspond to the case ws=0w_{s}=0. In this case [34, 33] the smooth component behaves exactly like a matter component scaling as ρs1/a3\rho_{s}\propto 1/a^{3}, so that both Ωm\Omega_{m} and Ωs\Omega_{s} are constant. They may thus be parameterized by this constant value of Ωm\Omega_{m} (or Ωs\Omega_{s}). In order to avoid confusion below with the general LCDM-type model where Ωm\Omega_{m} is a function of time, we will use, as in [34], the parameter

κ2=1Ωm.\kappa^{2}=\frac{1}{\Omega_{m}}\,. (3)

We then have that

2=8πG3a2κ2ρm,τ=4πG3a2κ2ρm,\begin{split}\mathcal{H}^{2}&=\frac{8\pi G}{3}a^{2}\kappa^{2}\rho_{m},\\ \frac{\partial{\mathcal{H}}}{\partial{\tau}}&=-\frac{4\pi G}{3}a^{2}\kappa^{2}\rho_{m},\end{split} (4)

and the scale factor a(t)a(t) evolves as that of an EdS model, with

a=(κtt0)2/3wheret0=16πGρm,0,\begin{split}a=\Big{(}\frac{\kappa t}{t_{0}}\Big{)}^{2/3}\quad\text{where}\quad t_{0}=\frac{1}{\sqrt{6\pi G\rho_{m,0}}}\,,\end{split} (5)

where ρm,0\rho_{m,0} is the matter density at some reference time. At the level of background cosmology this family of cosmologies are all equivalent to a standard EdS cosmology driven by a total “matter” density ρtot=κ2ρm\rho_{tot}=\kappa^{2}\rho_{m}. However, as soon as perturbations to uniformity are considered, the dependence on the evolution of the scale factor on κ2\kappa^{2} when expressed in terms of the mean (clustering) mass density translates into a physical difference of the models as a function of κ2\kappa^{2}. This is seen already, as we will show below, in the evolution of perturbations at linear order. As in the standard EdS case (κ=1\kappa=1) there is a growing mode and a decaying mode, but their temporal evolution is modified when κ1\kappa\neq 1. In particular the growing mode for density perturbations has the behaviour D(a)aαD(a)\propto a^{\alpha} where

α=14+141+24κ2.\alpha=-\frac{1}{4}+\frac{1}{4}\sqrt{1+\frac{24}{\kappa^{2}}}\,. (6)

As α\alpha is a one-to-one function of κ2\kappa^{2}, either can be used to characterize the one-parameter family of gEdS models. Because of the crucial role played in perturbation theory by the linear theory growing mode, we will often find it convenient to give our results as functions of α\alpha in our analysis below.

We note that while it is required that ρtot>0\rho_{tot}>0, and ρm>0\rho_{m}>0, it is not necessary that ρs>0\rho_{s}>0. Thus we consider the family of gEdS cosmologies to be parameterized either by κ[0,]\kappa\in[0,\infty] or, alternatively, by α[0,]\alpha\in[0,\infty]. Our motivation for studying these models here comes, as has been outlined in the introduction, from their interest as a theoretical tool rather than as relevant physical models. We note however that, for ρs>0\rho_{s}>0 (i.e. 0<Ωm<10<\Omega_{m}<1), a smooth component with zero pressure can model approximately in certain regimes a component in massive neutrinos, or also a contribution from the zero mode of a scalar field oscillating about the minimum of a quadratic potential or rolling in a simple exponential potential (see e.g. [37] and references therein). An alternative physical interpretation, which extends also to models with ρs<0\rho_{s}<0 (or κ2<1\kappa^{2}<1), is in terms of a model in which the gravitational coupling is scale-dependent, with the expansion being driven by a rescaled coupling κ2G\kappa^{2}G. The limit κ0\kappa\rightarrow 0 corresponds formally to the case of a static universe, in which a smooth component with negative energy exactly balances the clustering matter (for further detail, see [34]).

II.2 Eulerian Perturbation Theory (EPT)

Our starting point is the usual fluid equations for the evolution, under gravity only, of perturbations in the matter in an FLRW universe in the Newtonian (sub-horizon, non-relativistic) limit (see e.g. [14]):

δ(𝐱,τ)τ+{[1+δ(𝐱,τ)]𝐮(𝐱,τ)}\displaystyle\frac{\partial\delta(\mathbf{x},\tau)}{\partial\tau}+\nabla\cdot\big{\{}[1+\delta(\mathbf{x},\tau)]\mathbf{u}(\mathbf{x},\tau)\big{\}} =\displaystyle= 0,\displaystyle 0, (7)
𝐮(𝐱,τ)τ+(τ)𝐮(𝐱,τ)+[𝐮(𝐱,τ)]𝐮(𝐱,τ)\displaystyle\frac{\partial\mathbf{u}(\mathbf{x},\tau)}{\partial\tau}+\mathcal{H}(\tau)\mathbf{u}(\mathbf{x},\tau)+[\mathbf{u}(\mathbf{x},\tau)\cdot\nabla]\mathbf{u}(\mathbf{x},\tau) =\displaystyle= ϕ(𝐱,τ)1ρ(ρσij),\displaystyle-\nabla\phi(\mathbf{x},\tau)-\frac{1}{\rho}\nabla(\rho\sigma_{ij}), (8)
2ϕ(x,τ)\displaystyle\nabla^{2}\phi(\textbf{x},\tau) =\displaystyle= 32Ωm2δ,\displaystyle\frac{3}{2}\Omega_{m}\mathcal{H}^{2}\delta, (9)

where δ(𝐱,τ)ρ(x,τ)/ρ¯(τ)1\delta(\mathbf{x},\tau)\equiv\rho(x,\tau)/\bar{\rho}(\tau)-1 is the matter density fluctuation, with ρ(x,τ)\rho(x,\tau) the matter density and ρ¯(τ)\bar{\rho}(\tau) its mean value, 𝐮\mathbf{u} is the (peculiar) velocity, ϕ(𝐱,τ)\phi(\mathbf{x},\tau) is the (Newtonian) gravitational potential, and σij\sigma_{ij} is the velocity dispersion tensor. We will henceforth neglect this last term, treating the fluid in the single flow approximation.

Using the Fourier transform conventions

f(k)=d3rexp[ikr]f(r),f(r)=d3k(2π)3exp[ikr]f(k),\begin{split}f(\textbf{k})&=\int d^{3}r\exp\big{[}-i\textbf{k}\cdot\textbf{r}\big{]}f(\textbf{r}),\\ f(\textbf{r})&=\int\frac{d^{3}k}{(2\pi)^{3}}\exp\big{[}i\textbf{k}\cdot\textbf{r}\big{]}f(\textbf{k}),\end{split} (10)

Eq. (7) and the divergence of Eq. (8) can be combined to give

δ(k,τ)+θ(k,τ)=Sα~(k,τ),\delta^{\prime}(\textit{{k}},\tau)+\theta(\textit{{k}},\tau)=S_{\tilde{\alpha}}(\textit{{k}},\tau), (11)
θ(k,τ)+θ(k,τ)+32Ωm(a)2δ(k,τ)=Sβ~(k,τ),\theta^{\prime}(\textit{{k}},\tau)+\mathcal{H}\theta(\textit{{k}},\tau)+\frac{3}{2}\Omega_{m}(a)\mathcal{H}^{2}\delta(\textit{{k}},\tau)=S_{\tilde{\beta}}(\textit{{k}},\tau), (12)

where θ\theta is the divergence of the velocity field 𝐮\mathbf{u}, so that

u(k)=ikk2θ(k).\textbf{u}(\textbf{k})=-i\frac{\textbf{k}}{k^{2}}\theta(\textbf{k}). (13)

The velocity field can be assumed to have zero vorticity because the latter can be shown to always decay compared to the irrotational part (see e.g. [14]). The source terms SαS_{\alpha} and SβS_{\beta} in Eqs. (11) and (12) are given by

Sα~(k,τ)=d3q(2π)3α~(q,kq)θ(q,τ)δ(kq,τ),S_{\tilde{\alpha}}(\textit{{k}},\tau)=-\int\frac{\text{d}^{3}q}{(2\pi)^{3}}\tilde{\alpha}(\textit{{q}},\textit{{k}}-\textit{{q}})\theta(\textit{{q}},\tau)\delta(\textit{{k}}-\textit{{q}},\tau), (14)
Sβ~(k,τ)=d3q(2π)3β~(q,kq)θ(q,τ)θ(kq,τ),S_{\tilde{\beta}}(\textit{{k}},\tau)=-\int\frac{\text{d}^{3}q}{(2\pi)^{3}}\tilde{\beta}(\textit{{q}},\textit{{k}}-\textit{{q}})\theta(\textit{{q}},\tau)\theta(\textit{{k}}-\textit{{q}},\tau), (15)

where the coupling kernels α~\tilde{\alpha} and β~\tilde{\beta} respectively are defined as

α~(q1,q2)=q1.(q1+q2)q12,\tilde{\alpha}(\textit{{q}}_{1},\textit{{q}}_{2})=\frac{\textit{{q}}_{1}.(\textit{{q}}_{1}+\textit{{q}}_{2})}{\textit{q}_{1}^{2}}, (16)
β~(q1,q2)=12(q1+q2)2q1.q2q12q22.\tilde{\beta}(\textit{{q}}_{1},\textit{{q}}_{2})=\frac{1}{2}(\textit{{q}}_{1}+\textit{{q}}_{2})^{2}\frac{\textit{{q}}_{1}.\textit{{q}}_{2}}{\textit{q}_{1}^{2}\textit{q}_{2}^{2}}. (17)

For a gEdS cosmology, parametrized by κ2\kappa^{2}, combining Eqs. (11) and (12) with Eq. (9) to obtain a single second-order differential equation for the density fluctuations δ(x,τ)\delta(\textbf{x},\tau) and for the divergence of velocity fluctuations θ(x,τ)\theta(\textbf{x},\tau), respectively, we have

2{a2a232aa+32κ2}δ(k,a)\displaystyle\mathcal{H}^{2}\Big{\{}-a^{2}\partial_{a}^{2}-\frac{3}{2}a\partial_{a}+\frac{3}{2\kappa^{2}}\Big{\}}\delta(\textit{{k}},a) =\displaystyle= Sβ~(k,a)a(aSα~(k,a)),\displaystyle S_{\tilde{\beta}}(\textit{{k}},a)-\mathcal{H}\partial_{a}(aS_{\tilde{\alpha}}(\textit{{k}},a)), (18)
{a2a2+52aa+(1232κ2)}θ(k,a)\displaystyle\mathcal{H}\Big{\{}a^{2}\partial_{a}^{2}+\frac{5}{2}a\partial_{a}+\Big{(}\frac{1}{2}-\frac{3}{2\kappa^{2}}\Big{)}\Big{\}}\theta(\textit{{k}},a) =\displaystyle= a(aSβ~(k,a))32κ2Sα~(k,a),\displaystyle\partial_{a}(aS_{\tilde{\beta}}(\textit{{k}},a))-\frac{3}{2\kappa^{2}}\mathcal{H}S_{\tilde{\alpha}}(\textit{{k}},a), (19)

where a\partial_{a} is the partial derivative respect to scale factor aa.

We now proceed to solve these equations following the standard method (see e.g. [14]) in standard Eulerian perturbation theory (EPT). Working to linear order in δ\delta and θ\theta, Eq. (18) is just

a2a2δ(k,a)32aaδ(k,a)+32κ2δ(k,a)=0,-a^{2}\partial_{a}^{2}\delta(\textit{{k}},a)-\frac{3}{2}a\partial_{a}\delta(\textit{{k}},a)+\frac{3}{2\kappa^{2}}\delta(\textit{{k}},a)=0, (20)

which has linearly independent decaying and growing mode solutions so that

δ(k,a)=δ+(k,a)+δ(k,a)=D+(a)δ+,0(k)+D(a)δ,0(k),\begin{split}\delta(\textit{{k}},a)&=\delta_{+}(\textbf{k},a)+\delta_{-}(\textbf{k},a)\\ &=D_{+}(a)\delta_{+,0}(\textbf{k})+D_{-}(a)\delta_{-,0}(\textbf{k}),\end{split} (21)

where δ±,0(k)\delta_{\pm,0}(\textbf{k}) are constants fixed at some reference time, a=1a=1, and D±=aα±D_{\pm}=a^{\alpha_{\pm}} with

α±=14±141+24κ2.\alpha_{\pm}=-\frac{1}{4}\pm\frac{1}{4}\sqrt{1+\frac{24}{\kappa^{2}}}. (22)

We assume the growing mode initial conditions, corresponding to the solution at linear order which may be written

δ(k,a)=D(a)δ(1)(k),\delta(\textit{{k}},a)=D(a)\delta^{(1)}(\textbf{k}), (23)

where D(a)=aαD(a)=a^{\alpha}, and δ(1)(k)\delta^{(1)}(\textbf{k}) is the fluctuation at a=1a=1. At the same linear order we have the relation between density fluctuations and velocity

θ(k,τ)=δ(k,τ),\theta(\textit{{k}},\tau)=-\delta^{\prime}(\textit{{k}},\tau)\,, (24)

from which the solution at linear order for the velocity divergence follows,

θ(k,a)=(a)αD(a)δ(1)(k).\theta(\textit{{k}},a)=-\mathcal{H}(a)\alpha D(a)\delta^{(1)}(\textbf{k}). (25)

For the case κ=1\kappa=1 (and α=1\alpha=1) we recover the standard expression for the EdS model. These expressions are just special cases of the standard ones for FLRW cosmologies, which correspond to Eq. (23) and

θ(k,a)=(a)α1(a)D1(a)δ(1)(k),\theta(\textit{{k}},a)=-\mathcal{H}(a)\alpha_{1}(a)D_{1}(a)\delta^{(1)}(\textbf{k}), (26)

where D1(a)D_{1}(a) is the appropriate growth factor for the linear theory growing mode in the cosmology, and

α1(a)=dlnD1dlna.\alpha_{1}(a)=\frac{d\ln D_{1}}{d\ln a}. (27)

Taking leading order solutions as Eqs. (23) and (25), we now proceed to solve Eqs. (18) and (19) perturbatively to higher orders taking the ansatz

δ(k,a)=i=1Di(a)δ(i)(k),θ(k,a)=(a)αi=1Di(a)θ(i)(k),\begin{split}\delta(\textbf{k},a)&=\sum_{i=1}^{\infty}D^{i}(a)\,\delta^{(i)}(\textbf{k}),\\ \theta(\textbf{k},a)&=-\mathcal{H}(a)\alpha\sum_{i=1}^{\infty}D^{i}(a)\,\theta^{(i)}(\textbf{k}),\end{split} (28)

where Di(a)(aα)iD^{i}(a)\equiv(a^{\alpha})^{i}, and δ(i)\delta^{(i)}, θ(i)(k)\theta^{(i)}(\textbf{k}) are the contributions at order ii to each of these quantities. Such a “separable” ansatz means that all dependence on the cosmological model disappears when δ(k)\delta(\textbf{k}) and θ(k)\theta(\textbf{k}) are expressed in terms of their values at linear order. Such an ansatz solves the equations exactly in perturbation for the EdS model (see e.g. [14]). For the family of perturbed FLRW cosmologies with a perturbed matter component, one can consider the same ansatz with α\alpha replaced by α1(a)\alpha_{1}(a), and it has been shown [38, 39] that the condition for it be a solution is that the ratio

Ωmα12,\frac{\Omega_{m}}{\alpha_{1}^{2}}, (29)

is constant in time. While this condition is not satisfied in LCDM-like models, it is in gEdS models (in which both Ωm\Omega_{m} and α1\alpha_{1} are individually constant).

Using this ansatz (28) the n-th order solutions for the density and velocity fields can be expressed in exactly the same form as in standard EdS as

δ(n)(k)\displaystyle\delta^{(n)}(\textbf{k}) =\displaystyle= m=1n{d3qm(2π)3δ(1)(qm)}\displaystyle\prod_{m=1}^{n}\Big{\{}\int\frac{d^{3}q_{m}}{(2\pi)^{3}}\delta^{(1)}(\textbf{q}_{m})\Big{\}} (30)
×(2π)3δD(kq|1n)Fn(q1,,qn),\displaystyle\times(2\pi)^{3}\delta_{D}(\textbf{k}-\textbf{q}|_{1}^{n})F_{n}(\textbf{q}_{1},\dots,\textbf{q}_{n}),
θ(n)(k)\displaystyle\theta^{(n)}(\textbf{k}) =\displaystyle= m=1n{d3qm(2π)3δ(1)(qm)}\displaystyle\prod_{m=1}^{n}\Big{\{}\int\frac{d^{3}q_{m}}{(2\pi)^{3}}\delta^{(1)}(\textbf{q}_{m})\Big{\}} (31)
×(2π)3δD(kq|1n)Gn(q1,,qn).\displaystyle\times(2\pi)^{3}\delta_{D}(\textbf{k}-\textbf{q}|_{1}^{n})G_{n}(\textbf{q}_{1},\dots,\textbf{q}_{n}).

Working to second-order in the expansion, we have

2{a2a232aa+32κ2}a2αδ(2)(k)\displaystyle\mathcal{H}^{2}\Big{\{}-a^{2}\partial_{a}^{2}-\frac{3}{2}a\partial_{a}+\frac{3}{2\kappa^{2}}\Big{\}}a^{2\alpha}\delta^{(2)}(\textit{{k}}) (32)
=\displaystyle= d3q(2π)3β~(q,kq)δ(1)(q)δ(1)(kq)×[2α2a2α]\displaystyle-\int\frac{\text{d}^{3}q}{(2\pi)^{3}}\tilde{\beta}(\textit{{q}},\textit{{k}}-\textit{{q}})\delta^{(1)}(\textit{{q}})\delta^{(1)}(\textit{{k}}-\textit{{q}})\times\big{[}\mathcal{H}^{2}\alpha^{2}a^{2\alpha}\big{]}
d3q(2π)3α~(q,kq)δ(1)(q)×δ(1)(kq)\displaystyle-\int\frac{\text{d}^{3}q}{(2\pi)^{3}}\tilde{\alpha}(\textit{{q}},\textit{{k}}-\textit{{q}})\delta^{(1)}(\textit{{q}})\times\delta^{(1)}(\textit{{k}}-\textit{{q}})
×2[α2+2α2]a2α,\displaystyle\times\mathcal{H}^{2}\big{[}\frac{\alpha}{2}+2\alpha^{2}\big{]}a^{2\alpha},

from which we obtain the second-order density kernel F2(q1,q2)\textit{F}_{2}(\textit{{q}}_{1},\textit{{q}}_{2}) as

(3α2+α2)F2(q1,q2)=α2β~(q1,q2)+(α2+2α2)α~(q1,q2),\Big{(}3\alpha^{2}+\frac{\alpha}{2}\Big{)}\textit{F}_{2}(\textit{{q}}_{1},\textit{{q}}_{2})=\alpha^{2}\tilde{\beta}(\textit{{q}}_{1},\textit{{q}}_{2})+\Big{(}\frac{\alpha}{2}+2\alpha^{2}\Big{)}\tilde{\alpha}(\textit{{q}}_{1},\textit{{q}}_{2}), (33)

and thus finally

F2(q1,q2)=d2α~(q1,q2)+(1d2)β~(q1,q2),\textit{F}_{2}(\textit{{q}}_{1},\textit{{q}}_{2})=d_{2}\tilde{\alpha}(\textit{{q}}_{1},\textit{{q}}_{2})+(1-d_{2})\tilde{\beta}(\textit{{q}}_{1},\textit{{q}}_{2}), (34)

where

d2=(1+4α1+6α).d_{2}=\Big{(}\frac{1+4\alpha}{1+6\alpha}\Big{)}. (35)

For the velocity field divergence at the same order we have

{a2a2+52aa+(1232κ2)}[αa2αθ(2)(k)]\displaystyle\mathcal{H}\Big{\{}a^{2}\partial_{a}^{2}+\frac{5}{2}a\partial_{a}+(\frac{1}{2}-\frac{3}{2\kappa^{2}})\Big{\}}\big{[}-\mathcal{H}\alpha a^{2\alpha}\theta^{(2)}(\textit{{k}})\big{]} (36)
=\displaystyle= d3q(2π)3β~(q,kq)δ(1)(q)δ(1)(kq)\displaystyle-\int\frac{\text{d}^{3}q}{(2\pi)^{3}}\tilde{\beta}(\textit{{q}},\textit{{k}}-\textit{{q}})\delta^{(1)}(\textit{{q}})\delta^{(1)}(\textit{{k}}-\textit{{q}})
×[22α3a2α]d3q(2π)3α~(q,kq)\displaystyle\times\big{[}\mathcal{H}^{2}2\alpha^{3}a^{2\alpha}\big{]}-\int\frac{\text{d}^{3}q}{(2\pi)^{3}}\tilde{\alpha}(\textit{{q}},\textit{{k}}-\textit{{q}})
×δ(1)(q)δ(1)(kq)2[α22+α3]a2α,\displaystyle\times\delta^{(1)}(\textit{{q}})\delta^{(1)}(\textit{{k}}-\textit{{q}})\mathcal{H}^{2}\big{[}\frac{\alpha^{2}}{2}+\alpha^{3}\big{]}a^{2\alpha},

which leads to the solution

G2(q1,q2)=d~2α~(q1,q2)+(1d~2)β~(q1,q2)\textit{G}_{2}(\textit{{q}}_{1},\textit{{q}}_{2})=\tilde{d}_{2}\tilde{\alpha}(\textit{{q}}_{1},\textit{{q}}_{2})+(1-\tilde{d}_{2})\tilde{\beta}(\textit{{q}}_{1},\textit{{q}}_{2}) (37)

where

d~2=(1+2α1+6α).\tilde{d}_{2}=\Big{(}\frac{1+2\alpha}{1+6\alpha}\Big{)}. (38)

Continuing to third-order in the same manner we obtain

F3(q1,q2,q3)\displaystyle\textit{F}_{3}(\textit{{q}}_{1},\textit{{q}}_{2},\textit{{q}}_{3}) =\displaystyle= 12[d3α~(q1,q2+q3)F2(q2,q3)\displaystyle\frac{1}{2}\bigg{[}d_{3}\tilde{\alpha}(\textit{{q}}_{1},\textit{{q}}_{2}+\textit{{q}}_{3})\textit{F}_{2}(\textit{{q}}_{2},\textit{{q}}_{3}) (39)
+(1d3)β~(q1,q2+q3)G2(q2,q3)\displaystyle+(1-d_{3})\tilde{\beta}(\textit{{q}}_{1},\textit{{q}}_{2}+\textit{{q}}_{3})\textit{G}_{2}(\textit{{q}}_{2},\textit{{q}}_{3})
+[d3α~(q1+q2,q3)\displaystyle+\big{[}d_{3}\tilde{\alpha}(\textit{{q}}_{1}+\textit{{q}}_{2},\textit{{q}}_{3})
+(1d3)β~(q1+q2,q3)]\displaystyle+(1-d_{3})\tilde{\beta}(\textit{{q}}_{1}+\textit{{q}}_{2},\textit{{q}}_{3})\big{]}
×G2(q1,q2)],\displaystyle\times\textit{G}_{2}(\textit{{q}}_{1},\textit{{q}}_{2})\bigg{]},

and

G3(q1,q2,q3)\displaystyle\textit{G}_{3}(\textit{{q}}_{1},\textit{{q}}_{2},\textit{{q}}_{3}) =\displaystyle= 12[d~3α~(q1,q2+q3)F2(q2,q3)\displaystyle\frac{1}{2}\bigg{[}\tilde{d}_{3}\tilde{\alpha}(\textit{{q}}_{1},\textit{{q}}_{2}+\textit{{q}}_{3})\textit{F}_{2}(\textit{{q}}_{2},\textit{{q}}_{3}) (40)
+(1d~3)β~(q1,q2+q3)G2(q2,q3)\displaystyle+(1-\tilde{d}_{3})\tilde{\beta}(\textit{{q}}_{1},\textit{{q}}_{2}+\textit{{q}}_{3})\textit{G}_{2}(\textit{{q}}_{2},\textit{{q}}_{3})
+[d~3α~(q1+q2,q3)\displaystyle+\big{[}\tilde{d}_{3}\tilde{\alpha}(\textit{{q}}_{1}+\textit{{q}}_{2},\textit{{q}}_{3})
+(1d~3)β~(q1+q2,q3)]\displaystyle+(1-\tilde{d}_{3})\tilde{\beta}(\textit{{q}}_{1}+\textit{{q}}_{2},\textit{{q}}_{3})\big{]}
×G2(q1,q2)],\displaystyle\times\textit{G}_{2}(\textit{{q}}_{1},\textit{{q}}_{2})\bigg{]},

where

d3=1+6α1+8α,d~3=1+2α1+8α.d_{3}=\frac{1+6\alpha}{1+8\alpha},\quad\tilde{d}_{3}=\frac{1+2\alpha}{1+8\alpha}. (41)

Higher-order kernels can be inferred by continuing this recursion.111Further details of these calculations are given in Appendix A.

Refer to caption
Figure 1: Variation of the functions d2d_{2} and d3d_{3} determining the EPT kernels at second and third order relative to their values in EdS, as a function of the logarithmic growth rate α\alpha of gEdS. The dashed horizontal lines bracket approximately the region in which the effective growth rate α1(a)\alpha_{1}(a) varies in LCDM-like models.

Fig. 1 shows how the kernels F2F_{2} and G2G_{2} vary relative to those in the standard EdS case as a function of α\alpha (and F3F_{3} and G3G_{3} have very similar behaviours). We see that the dependence on α\alpha is very weak, other than at values of α\alpha very much less than unity. Indeed expanding about α=1\alpha=1 we have

d2=d2(1)[1+235(1α)+O((1α)2)],\displaystyle d_{2}=d_{2}(1)\left[1+\frac{2}{35}(1-\alpha)+O((1-\alpha)^{2})\right],
d3=d3(1)[1+263(1α)+O((1α)2)],\displaystyle d_{3}=d_{3}(1)\left[1+\frac{2}{63}(1-\alpha)+O((1-\alpha)^{2})\right], (42)

and, over the whole range of α\alpha, varying from 0 to \infty, we have

75>d2d2(1)>1415,\displaystyle\frac{7}{5}>\frac{d_{2}}{d_{2}(1)}>\frac{14}{15}, (43)
97>d3d3(1)>2728.\displaystyle\frac{9}{7}>\frac{d_{3}}{d_{3}(1)}>\frac{27}{28}. (44)

A naive guess would be that, in a generic FLRW cosmology, the kernel at any time can be approximated by those in a gEdS model, with the growth rate corresponding to the instantaneous value of its effective logarithmic growth rate α1(a)\alpha_{1}(a). For standard-type cosmological models, as we will discuss in detail in the next section, such an effective growth rate decreases relative to EdS to at most α0.5\alpha\approx 0.5 at z=0z=0 as dark energy contributions accelerate the expansion. From Fig. 1 we see that this would correspond to a change of about 5%5\% (3%3\%) in d2d_{2}(d3d_{3}) relative to EdS. In practice, as we will see in the second part of the paper, such a mapping to an approximate mapping may indeed to used, but the relevant effective growth exponent is given by this naive guess only for a sufficiently slowly varying smooth component. For LCDM-like models it is much closer to the initial value. The very small sub-percent corrections calculated for LCDM-like models can thus be understood as a combination of the very weak sensitivity of the gEdS kernels to the growth rate of fluctuations, and the rapid evolution of dark energy at low redshift.

II.3 Lagrangian Perturbation Theory (LPT)

In LPT one characterizes the evolution of the fluid using its displacement field Ψ(q,t)\Psi(\textbf{q},t) as a function of its initial Lagrangian position q of the particle, in terms of which the Eulerian position is given by

x(q,τ)=q+𝚿(q,τ).\textbf{x}(\textbf{q},\tau)=\textbf{q}+\mathbf{\Psi}(\textbf{q},\tau). (45)

It follows in particular that

[1+δ(x)]d3x=d3q,\big{[}1+\delta(\textbf{x})\big{]}\text{d}^{3}\text{x}=\text{d}^{3}\text{q}, (46)

and that the determinant JJ of the Jacobian matrix is given by

J=|d3xd3q|=det[δij(K)+Ψi,j],J=\Big{|}\frac{d^{3}x}{d^{3}q}\Big{|}=\text{det}\Big{[}\delta_{ij}^{(K)}+\Psi_{i,j}\Big{]}, (47)

where

Ψi,j=Ψiqj,\Psi_{i,j}=\frac{\partial\Psi_{i}}{\partial q_{j}}, (48)

so that

δ(x)=1J1.\delta(\textbf{x})=\frac{1}{J}-1. (49)

The equation of motion for the fluid particle is

d2𝚿dτ2+d𝚿dτ=xϕ.\frac{\text{d}^{2}\mathbf{\Psi}}{\text{d}\tau^{2}}+\mathcal{H}\frac{\text{d}\mathbf{\Psi}}{\text{d}\tau}=-\nabla_{x}\phi. (50)

Taking the divergence of this equation, and using the Poisson Eq. (9), one obtains

Jx[d2𝚿dτ2+d𝚿dτ]=32Ωm2[J1].\begin{split}J\bm{\nabla}_{\textbf{x}}\cdot\Big{[}\frac{\text{d}^{2}\mathbf{\Psi}}{\text{d}\tau^{2}}+\mathcal{H}\frac{\text{d}\mathbf{\Psi}}{\text{d}\tau}\Big{]}&=\frac{3}{2}\Omega_{m}\mathcal{H}^{2}[J-1].\end{split} (51)

This equation order is then solved order by order using the perturbative expansion for

𝚿=𝚿(1)+𝚿(2)+𝚿(3)+\mathbf{\Psi}=\mathbf{\Psi}^{(1)}+\mathbf{\Psi}^{(2)}+\mathbf{\Psi}^{(3)}+\dots (52)

and

J=1+J(1)+J(2)+J(3)+J=1+J^{(1)}+J^{(2)}+J^{(3)}+\dots (53)

where the latter are related to the former up to third order by the following expressions (for a detailed derivation, see e.g. [40]):

J(1)=iΨi,i(1),J(2)=iΨi,i(2)+12ij{Ψi,i(1)Ψj,j(1)Ψi,j(1)Ψj,i(1)},J(3)=iΨi,i(3)+ij{Ψi,i(2)Ψj,j(1)Ψi,j(2)Ψj,i(1)}+detΨi,j(1).\begin{split}J^{(1)}&=\sum_{i}\Psi_{i,i}^{(1)},\\ J^{(2)}&=\sum_{i}\Psi_{i,i}^{(2)}+\frac{1}{2}\sum_{i\neq j}\Big{\{}\Psi_{i,i}^{(1)}\Psi_{j,j}^{(1)}-\Psi_{i,j}^{(1)}\Psi_{j,i}^{(1)}\Big{\}},\\ J^{(3)}&=\sum_{i}\Psi_{i,i}^{(3)}+\sum_{i\neq j}\Big{\{}\Psi_{i,i}^{(2)}\Psi_{j,j}^{(1)}-\Psi_{i,j}^{(2)}\Psi_{j,i}^{(1)}\Big{\}}+\text{det}\Psi_{i,j}^{(1)}.\end{split} (54)

Using the chain rule

xi=(δij(K)+Ψi,j)1qj,\nabla_{x_{i}}=(\delta_{ij}^{(K)}+\Psi_{i,j})^{-1}\nabla_{q_{j}}, (55)

one then solves Eq. (51) order by order. At linear order we have simply

d2Ψi,i(1)dτ2+dΨi,i(1)dτ32Ωm2Ψi,i(1)=0,\frac{\text{d}^{2}\Psi_{i,i}^{(1)}}{\text{d}\tau^{2}}+\mathcal{H}\frac{\text{d}\Psi_{i,i}^{(1)}}{\text{d}\tau}-\frac{3}{2}\Omega_{m}\mathcal{H}^{2}\Psi_{i,i}^{(1)}=0, (56)

and thus, choosing the growing mode solution,

Ψi,i(1)(x,τ)=δ(1)(x,τ),\Psi_{i,i}^{(1)}(\textbf{x},\tau)=-\delta^{(1)}(\textbf{x},\tau), (57)

which gives in Fourier space

Ψ(1)(k,τ)=ikk2δ(1)(k)D1(τ).\Psi^{(1)}(\textbf{k},\tau)=-\frac{i\textbf{k}}{k^{2}}\delta^{(1)}(\textbf{k})D_{1}(\tau). (58)

This is the solution for any perturbed FLRW cosmology and thus for gEdS in particular with D1(a)=D(a)=aαD_{1}(a)=D(a)=a^{\alpha}.

Considering now gEdS cosmologies, we proceed to solve Eq. (51) to non-linear order making the separable ansatz analogous to that in EPT:

Ψ(n)(k,τ)=Dn(τ)Ψ(n)(k).\Psi^{(n)}(\textbf{k},\tau)=D^{n}(\tau)\Psi^{(n)}(\textbf{k})\,. (59)

We have then

(1+J(1)+J(2)+J(3)+)(δij(K)Ψi,j+Ψi,lΨl,j+)×nα[nα+12]Ψi,j=32κ2[J(1)+J(2)+J(3)+],\begin{split}&(1+J^{(1)}+J^{(2)}+J^{(3)}+\dots)(\delta_{ij}^{(K)}-\Psi_{i,j}+\Psi_{i,l}\Psi_{l,j}+\dots)\\ &\times n\alpha\Big{[}n\alpha+\frac{1}{2}\Big{]}\Psi_{i,j}=\frac{3}{2\kappa^{2}}[J^{(1)}+J^{(2)}+J^{(3)}+\dots],\end{split} (60)

where we have used dda=2a\frac{\text{d}\mathcal{H}}{\text{d}a}=-\frac{\mathcal{H}}{2a}. At second order we have therefore

Ψi,i(2)=2α+12(6α+1)i,j[Ψi,i(1)Ψj,j(1)Ψi,j(1)Ψi,j(1)].\Psi_{i,i}^{(2)}=-\frac{2\alpha+1}{2(6\alpha+1)}\sum_{i,j}\big{[}\Psi_{i,i}^{(1)}\Psi_{j,j}^{(1)}-\Psi_{i,j}^{(1)}\Psi_{i,j}^{(1)}\Big{]}. (61)

The result at third-order, of which further details are given in Appendix A.2, is

Ψi,i(3)\displaystyle\Psi_{i,i}^{(3)} =\displaystyle= 1(8α+1)[(4α+1){Ψi,i(2)Ψj,j(1)+Ψi,j(2)Ψj,i(1)}\displaystyle\frac{1}{(8\alpha+1)}\Big{[}(4\alpha+1)\Big{\{}-\Psi_{i,i}^{(2)}\Psi_{j,j}^{(1)}+\Psi_{i,j}^{(2)}\Psi_{j,i}^{(1)}\Big{\}} (62)
+(2α+1){12Ψi,i(1)Ψi,j(1)Ψj,i(1)\displaystyle+(2\alpha+1)\Big{\{}\frac{1}{2}\Psi_{i,i}^{(1)}\Psi_{i,j}^{(1)}\Psi_{j,i}^{(1)}
16Ψi,i(1)Ψj,j(1)Ψk,k(1)13Ψi,k(1)Ψk,j(1)Ψj,i(1)}].\displaystyle-\frac{1}{6}\Psi_{i,i}^{(1)}\Psi_{j,j}^{(1)}\Psi_{k,k}^{(1)}-\frac{1}{3}\Psi_{i,k}^{(1)}\Psi_{k,j}^{(1)}\Psi_{j,i}^{(1)}\Big{\}}\Big{]}.

At n-th order the displacement field in Fourier space can be written in the form

Ψ(n)(k)\displaystyle\Psi^{(n)}(\textbf{k}) =\displaystyle= ini=1n{d3qi(2π)3δ(1)(qi)}Ln(q1,,qn)\displaystyle-\frac{i}{n}\prod_{i=1}^{n}\Big{\{}\int\frac{\text{d}^{3}q_{i}}{(2\pi)^{3}}\delta^{(1)}(\textbf{q}_{i})\Big{\}}\textbf{L}_{n}(\textbf{q}_{1},\dots,\textbf{q}_{n}) (63)
×(2π)3δ(D)(kq1qn),\displaystyle\times(2\pi)^{3}\delta^{(D)}(\textbf{k}-\textbf{q}_{1}\dots\textbf{q}_{n}),

where the first, second, and third-order solutions correspond to the following kernels:

L1\displaystyle\textbf{L}_{1} =\displaystyle= kk2,\displaystyle\frac{\textbf{k}}{k^{2}}, (64)
L2\displaystyle\textbf{L}_{2} =\displaystyle= (2α+16α+1)kk2[1(q1q2)2q12q22],\displaystyle\Big{(}\frac{2\alpha+1}{6\alpha+1}\Big{)}\frac{\textbf{k}}{k^{2}}\Big{[}1-\frac{(\textbf{q}_{1}\cdot\textbf{q}_{2})^{2}}{q_{1}^{2}q_{2}^{2}}\Big{]}, (65)
L3\displaystyle\textbf{L}_{3} =\displaystyle= 3(4α+18α+1)(2α+16α+1)kk2[1(q1q2q1q2)2]\displaystyle 3\Big{(}\frac{4\alpha+1}{8\alpha+1}\Big{)}\Big{(}\frac{2\alpha+1}{6\alpha+1}\Big{)}\frac{\textbf{k}}{k^{2}}\Big{[}1-\Big{(}\frac{\textbf{q}_{1}\cdot\textbf{q}_{2}}{q_{1}q_{2}}\Big{)}^{2}\Big{]} (66)
×[1((q1+q2)q3|(q1+q2|q3)2](2α+18α+1)kk2\displaystyle\times\Big{[}1-\Big{(}\frac{(\textbf{q}_{1}+\textbf{q}_{2})\cdot\textbf{q}_{3}}{|(\textbf{q}_{1}+\textbf{q}_{2}|q_{3}}\Big{)}^{2}\Big{]}-\Big{(}\frac{2\alpha+1}{8\alpha+1}\Big{)}\frac{\textbf{k}}{k^{2}}
×[13(q1q2q1q2)2\displaystyle\times\Big{[}1-3\Big{(}\frac{\textbf{q}_{1}\cdot\textbf{q}_{2}}{q_{1}q_{2}}\Big{)}^{2}
+2(q1q2)(q2q3)(q3q1)q12q22q32].\displaystyle+2\frac{(\textbf{q}_{1}\cdot\textbf{q}_{2})(\textbf{q}_{2}\cdot\textbf{q}_{3})(\textbf{q}_{3}\cdot\textbf{q}_{1})}{q_{1}^{2}q_{2}^{2}q_{3}^{2}}\Big{]}.

As a check on our calculation we can use the known relations between the EPT and LPT kernels (see [29, 41, 42]). At first-order we have simply

F1(k)=kL1(k)=kkk2=1.F_{1}(\textbf{k})=\textbf{k}\cdot\textbf{L}_{1}(\textbf{k})=\textbf{k}\cdot\frac{\textbf{k}}{k^{2}}=1. (67)

The second-order density kernels are related by

F2(q1,q2)\displaystyle F_{2}(\textbf{q}_{1},\textbf{q}_{2}) =\displaystyle= 12[kL2(q1,q2)+[kL1(q1][kL1(q2)]]\displaystyle\frac{1}{2}\Big{[}\textbf{k}\cdot\textbf{L}_{2}(\textbf{q}_{1},\textbf{q}_{2})+[\textbf{k}\cdot\textbf{L}_{1}(\textbf{q}_{1}][\textbf{k}\cdot\textbf{L}_{1}(\textbf{q}_{2})]\Big{]} (68)
=\displaystyle= 4α+16α+1+12q1q2q1q2[q2q1+q1q2]\displaystyle\frac{4\alpha+1}{6\alpha+1}+\frac{1}{2}\frac{\textit{{q}}_{1}\cdot\textit{{q}}_{2}}{q_{1}q_{2}}\Big{[}\frac{q_{2}}{q_{1}}+\frac{q_{1}}{q_{2}}\Big{]}
+(2α6α+1)(q1q2)2q12q22,\displaystyle+\Big{(}\frac{2\alpha}{6\alpha+1}\Big{)}\frac{(\textit{{q}}_{1}\cdot\textit{{q}}_{2})^{2}}{q_{1}^{2}q_{2}^{2}},

while at third order we have

F3(q1,q2,q3)\displaystyle F_{3}(\textbf{q}_{1},\textbf{q}_{2},\textbf{q}_{3}) =\displaystyle= 13![kL3(s)(q1,q2,q3)\displaystyle\frac{1}{3!}\Big{[}\textbf{k}\cdot\textbf{L}_{3}^{(s)}(\textbf{q}_{1},\textbf{q}_{2},\textbf{q}_{3}) (69)
+{[kL1(q1)][kL2(q2,q3)]\displaystyle+\big{\{}[\textbf{k}\cdot\textbf{L}_{1}(\textbf{q}_{1})][\textbf{k}\cdot\textbf{L}_{2}(\textbf{q}_{2},\textbf{q}_{3})]
+cyc}\displaystyle+\text{cyc}\big{\}}
+[kL1(q1)][kL1(q2)]\displaystyle+[\textbf{k}\cdot\textbf{L}_{1}(\textbf{q}_{1})][\textbf{k}\cdot\textbf{L}_{1}(\textbf{q}_{2})]
×[kL1(q3)]].\displaystyle\times[\textbf{k}\cdot\textbf{L}_{1}(\textbf{q}_{3})]\Big{]}.

It is straightforward to check that we indeed recover the results in EPT of the previous subsection.

III Power spectrum in gEdS models

We define the power spectrum P(k)P(k)P(\vec{k})\equiv P(k) (k=|k|k=|\textbf{k}|) of the (assumed) statistically homogeneous and isotropic stochastic density field by

δ(k,a)δ(k,a)=(2π)3δ(D)(k+k)P(|k|,a),\langle\delta(\textbf{k},a)\delta(\textbf{k}^{\prime},a)\rangle=(2\pi)^{3}\delta^{(D)}(\textbf{k}+\textbf{k}^{\prime})P(|\textbf{k}|,a), (70)

where \langle\cdots\rangle denotes the ensemble average. Assuming that the fluctuations are Gaussian at linear order, one obtains, retaining terms up to fourth order in the linear order solution,

δ(k,a)δ(k,a)\displaystyle\langle\delta(\textbf{k},a)\delta(\textbf{k}^{\prime},a)\rangle =\displaystyle= δ(1)(k,a)δ(1)(k,a)\displaystyle\langle\delta^{(1)}(\textbf{k},a)\delta^{(1)}(\textbf{k}^{\prime},a)\rangle (71)
+2δ(1)(k,a)δ(3)(k,a)\displaystyle+2\langle\delta^{(1)}(\textbf{k},a)\delta^{(3)}(\textbf{k}^{\prime},a)\rangle
+δ(2)(k,a)δ(2)(k,a),\displaystyle+\langle\delta^{(2)}(\textbf{k},a)\delta^{(2)}(\textbf{k}^{\prime},a)\rangle,

and thus the expression for the “one-loop” power spectrum as

P1loop(k,a)=Plin(k,a)+P22(k,a)+2P13(k,a),\textit{P}_{1-\text{loop}}(k,a)=P_{\text{\rm lin}}(k,a)+P_{22}(k,a)+2P_{13}(k,a), (72)

where Plin(k,a)P_{\text{lin}}(k,a) is the linear power spectrum (i.e. defined by Eq. (70) with δ(k,a)=D(a)δ(1)(k\delta(\vec{k},a)=D(a)\delta^{(1)}(\vec{k}), and the two other terms correspond to

P22(k,a)\displaystyle P_{22}(k,a) =\displaystyle= 2d3q(2π)3Plin(q,a)Plin(|kq|,a)\displaystyle 2\int\frac{d^{3}q}{(2\pi)^{3}}P_{\text{lin}}(q,a)P_{\text{lin}}(|\textbf{k}-\textbf{q}|,a) (73)
×|F2(s)(kq,q)|2.\displaystyle\times|F_{2}^{(s)}(\textbf{k}-\textbf{q},\textbf{q})|^{2}.

and

P13(k,a)\displaystyle P_{13}(k,a) =\displaystyle= 3Plin(k,a)d3q(2π)3Plin(q,a)\displaystyle 3P_{\text{lin}}(k,a)\int\frac{d^{3}q}{(2\pi)^{3}}P_{\text{lin}}(q,a) (74)
×F3(s)(k,q,q),\displaystyle\times F_{3}^{(s)}(\textbf{k},\textbf{q},-\textbf{q}),

where F2(s)F_{2}^{(s)} and F3(s)F_{3}^{(s)} are the symmetrized form of F2F_{2} and F3F_{3} (with respect to permutation of their arguments). These expressions are formally identical to those in standard EdS, and share the property of any separable solution that the perturbative corrections to the PS at any time are functionals only of the linear power spectrum at that time. The modifications of gEdS relative to the standard EdS model are expressed solely through the α\alpha-dependence of the kernels F2F_{2} and F3F_{3} which we have derived above. Using these to make the α\alpha-dependence explicit we obtain

P22\displaystyle P_{22} =\displaystyle= M0+1+4α1+6αM1+(1+4α1+6α)2M2,\displaystyle M_{0}+\frac{1+4\alpha}{1+6\alpha}M_{1}+\left(\frac{1+4\alpha}{1+6\alpha}\right)^{2}M_{2}, (75)
2P13\displaystyle 2P_{13} =\displaystyle= N0+1+2α1+8αN1+2α(1+2α)(1+6α)(1+8α)N2,\displaystyle N_{0}+\frac{1+2\alpha}{1+8\alpha}N_{1}+\frac{2\alpha(1+2\alpha)}{(1+6\alpha)(1+8\alpha)}N_{2}, (76)

where the MiM_{i} are the integrals

Mi(k,a)\displaystyle M_{i}(k,a) =\displaystyle= 18π2k30𝑑rPlin(kr,a)\displaystyle\frac{1}{8\pi^{2}}k^{3}\int_{0}^{\infty}drP_{\text{lin}}(kr,a) (77)
×11dμPlin(k1+r22μr,a)(1+r22μr)2\displaystyle\times\int_{-1}^{1}d\mu\frac{P_{\text{lin}}(k\sqrt{1+r^{2}-2\mu r},a)}{(1+r^{2}-2\mu r)^{2}}\,
×mi(r,μ),\displaystyle\times m_{i}(r,\mu),

with

m0(r,μ)\displaystyle m_{0}(r,\mu) =\displaystyle= (μr)2,\displaystyle(\mu-r)^{2}, (78)
m1(r,μ)\displaystyle m_{1}(r,\mu) =\displaystyle= 4r(μr)(1μ2),\displaystyle 4r(\mu-r)(1-\mu^{2}), (79)
m2(r,μ)\displaystyle m_{2}(r,\mu) =\displaystyle= 4r2(1μ2)2,\displaystyle 4r^{2}(1-\mu^{2})^{2}, (80)

and NiN_{i} are the integrals

Ni(k,a)=18π2k3Plin(k,a)0𝑑rPlin(kr,a)ni(r)N_{i}(k,a)=\frac{1}{8\pi^{2}}k^{3}P_{\rm lin}(k,a)\int_{0}^{\infty}drP_{\rm lin}(kr,a)n_{i}(r) (81)

with

n0(r)=43,n1(r)=1+83r2r4+(r21)32rln|1+r||1r|,n2(r)=1r2(183r2r4)+(r21)32r3ln|1+r||1r|,\begin{split}n_{0}(r)&=-\frac{4}{3},\\ n_{1}(r)&=1+\frac{8}{3}r^{2}-r^{4}+\frac{(r^{2}-1)^{3}}{2r}\ln\frac{|1+r|}{|1-r|},\\ n_{2}(r)&=\frac{1}{r^{2}}(1-\frac{8}{3}r^{2}-r^{4})+\frac{(r^{2}-1)^{3}}{2r^{3}}\ln\frac{|1+r|}{|1-r|},\end{split} (82)

and we have defined μ=kq/kq\mu=\textbf{k}\cdot\textbf{q}/kq and r=q/kr=q/k.

III.1 Convergence properties of the one-loop power spectrum

IR-divergent UV-divergent
M0M_{0} n1n\leq-1 n1/2n\geq 1/2
M1M_{1} n3n\leq-3 n1/2n\geq 1/2
M2M_{2} n3n\leq-3 n1/2n\geq 1/2
N0N_{0} n1n\leq-1 n1n\geq-1
N1N_{1} n3n\leq-3 n1n\geq-1
N2N_{2} n3n\leq-3 n1n\geq-1
N0+M0N_{0}+M_{0} n3n\leq-3 n1n\geq-1
Table 1: Convergence properties of the integrals MiM_{i} and NiN_{i} (i=0,1,2i=0,1,2). In each case the bound on nn is that obtained assuming that PlinknP_{\rm lin}\sim k^{n} in the relevant (k0k\rightarrow 0 or kk\rightarrow\infty) limit. Integrals are “infra-red safe” if they converge for n>3n>-3, i.e. when Plin(k)P_{\rm lin}(k) itself is integrable (in three dimensions) as k0k\rightarrow 0. The cancellation of the divergences for 1n>3-1\geq n>-3 in the sum N0+M0N_{0}+M_{0} corresponds to the well-known cancellation between the full P22P_{22} and P13P_{13} contributions in an EdS cosmology. As expected this cancellation generalizes to the gEdS case, but without the need for any cancellation in the α\alpha-dependent terms. We will show below that the corrections in standard (e.g. LCDM-like) cosmologies to the one-loop PS are expressed in terms of the four infra-red safe integrals M1,M2,N1,N2M_{1},M_{2},N_{1},N_{2}.

We now consider the convergence properties of these one-loop contributions to the PS in the gEdS model i.e. for what asymptotic behaviours of the linear PS, Plin(k,a)P_{\text{lin}}(k,a), the integrals converge. The results of this analysis are summarized in Table 1. These are obtained straightforwardly by considering the r0r\rightarrow 0, and rr\rightarrow\infty, behaviours of the integrands for the infra-red and ultra-violet limits, respectively. In the MiM_{i} integrals there is also a divergence in the angular integral at μ=1\mu=1 for r=1r=1, corresponding to |𝐪𝐤|𝟎|\bf{q}-\bf{k}|\rightarrow 0, but because of the symmetry of exchange of the variables 𝐤\bf{k} and 𝐪𝐤\bf{q}-\bf{k} in the integral (as noted e.g. in [13]), its contribution is identical to that from r0r\rightarrow 0 and is thus taken into account by multiplying it by a factor of two.

The leading contribution to M0M_{0} as r0r\rightarrow 0 corresponds to m0=μ2m_{0}=\mu^{2}, which when integrated over angle, and multiplied by two, gives a factor of 4/34/3 which can be seen to exactly cancel the corresponding leading contribution from N0N_{0}. These contributions to each integral are proportional to Plin(k)𝑑k\int P_{\rm lin}(k)dk i.e. to the variance of the displacement field. As understood in early works on perturbation theory (e.g. [6, 43], see also [44] for a useful discussion), these are unphysical divergences if the system (as is the case here, just as in standard EdS) is Galilean invariant, and their cancellation is a consequence of this invariance. The next term in the expansion as r0r\rightarrow 0 does not cancel but is proportional to Plin(k)k2𝑑k\int P_{\rm lin}(k)k^{2}dk i.e. to the variance of the fluctuation field (which is Galilean invariant). Their sum M0+N0M_{0}+N_{0} is then said to be “infra-red safe” in this case: the integral converges for any Plin(k,a)P_{\text{lin}}(k,a) which is integrable at k0k\rightarrow 0.

We see, on the other hand, that the integrals M1,M2,N1M_{1},M_{2},N_{1} and N2N_{2} are all infra-red safe: for r0r\rightarrow 0, the functions m1,m2,n1,n2m_{1},m_{2},n_{1},n_{2} all have, after integration over angles, a leading behaviour r2\sim r^{2}. This means that infra-red safety of all the α\alpha-dependent contributions is obtained term by term, without any cancellation between different terms. This is evidently a necessary condition to obtain infra-red safety given that the α\alpha-dependence in the F2F_{2} and F3F_{3} kernels are given through independent non-linear functions of α\alpha. As we will see below it turns out that we can write the cosmological corrections (due to departures from EdS) of the one-loop PS in LCDM-like models in terms of the same integrals. Infra-red safety of these cosmological corrections is thus explicit and their numerical calculation simplified, without the additional manipulations needed for the leading EdS approximated contribution (see e.g. [45]).

Although we will not discuss them further in this paper, it is interesting to consider also the ultraviolet divergences. These are physical divergences marking a real breakdown of standard perturbation theory. It is straightforward to show (see Appendix B for details) that combining the leading contributions in the six integrals above we obtain

P22(k,a)=7+36α+92α230(1+6α)2k4d3q(2π)3Plin2(q)q4+,P_{22}(k,a)=\frac{7+36\alpha+92\alpha^{2}}{30(1+6\alpha)^{2}}k^{4}\int\frac{d^{3}q}{(2\pi)^{3}}\frac{P_{\text{lin}}^{2}(q)}{q^{4}}+\cdots, (83)

and

2P13(k)\displaystyle 2P_{13}(k) =\displaystyle= 714α176α215(1+6α)(1+8α)k2Plin(k)d3q(2π)3Plin(q)q2\displaystyle\frac{7-14\alpha-176\alpha^{2}}{15(1+6\alpha)(1+8\alpha)}k^{2}P_{\text{lin}}(k)\int\frac{d^{3}q}{(2\pi)^{3}}\frac{P_{\text{lin}}(q)}{q^{2}} (84)
+12(8α1)(1+2α)105(1+6α)(1+8α)k4Plin(k)\displaystyle+\frac{12(8\alpha-1)(1+2\alpha)}{105(1+6\alpha)(1+8\alpha)}k^{4}P_{\text{lin}}(k)
×d3q(2π)3Plin(q)q4+\displaystyle\times\int\frac{d^{3}q}{(2\pi)^{3}}\frac{P_{\text{lin}}(q)}{q^{4}}+\cdots

where the dots now indicate terms proportional to integrals which converge more rapidly in the ultra-violet than these leading terms. The first expression corresponds to the sum of the leading contribution in the integrals MiM_{i}, which we see (as indicated in Table 1) diverge for n1/2n\geq 1/2, while the second shows the sum of the leading contributions from the NiN_{i}, the first diverging for n1n\geq-1 (as indicated in Table 1) and the next one diverging for n1n\geq 1.

The integrals in these expressions are the same as those in the standard EdS models, the only difference being in the coefficients which are manifestly all explicitly α\alpha dependent. The leading divergence overall is that in P13P_{13}, diverging for n1n\geq-1. We note that this remains true in the family of gEdS models, except that there is a specific value, α=0.1635\alpha=0.1635\cdots, at which the coefficient of this leading term, proportional to 714α176α27-14\alpha-176\alpha^{2}, vanishes. For this specific value of α\alpha therefore the one-loop PS is ultra-violet convergent for n<1/2n<1/2. We will not explore further in this article the physical significance of this result, nor more generally the α\alpha dependence of the ultra-violet divergences. To address these questions it is necessary to consider the regularisation of these divergences, using for example, the RPT or EFT approaches (see references cited in the introduction above). We will explore these issues in future work.

IV LCDM approximated as an interpolation of gEdS

In the rest of this article we explore the relation of these models to standard (e.g. LCDM-like) cosmological models. The class of models considered are the FLRW models described at the beginning of Section II.1, i.e., with a clustering matter component and a smooth component with a generic equation of state. Further to this we will assume only that cosmology is matter dominated at early times, i.e. that it is EdS, or gEdS if there remains a smooth matter-like component (e.g. massive neutrinos or matter-like dark energy). Specifically we explore whether we can make use of the analytical results for perturbation theory in gEdS to calculate, exactly or approximately, results in standard models (see references further below). More precisely we wish to consider whether the functions, Q(z)Q(z) say, characterising the evolution of fluctuations in perturbation theory in an LCDM model as a function of redshift zz may be interpolated on the family of gEds models, so that we can write

Q(z)QgEdS(α=αeff(z)),Q(z)\approx Q_{gEdS}(\alpha=\alpha_{eff}(z)), (85)

where αeff(z)\alpha_{eff}(z) is an effective growth exponent which can be calculated given the parameters of the LCDM model. Further we will determine whether the approximation

Q(z)QgEdS(κ2=1/Ωm(z)),Q(z)\approx Q_{gEdS}(\kappa^{2}=1/\Omega_{m}(z)), (86)

is valid, i.e., whether we can take

αeff(z)=14+141+24Ωm(z),\alpha_{eff}(z)=-\frac{1}{4}+\frac{1}{4}\sqrt{1+24\Omega_{m}(z)}, (87)

which means that we match both the instantaneous expansion rate and growth rate of the FLRW cosmology to that of a gEdS cosmology. We will say in this case that the interpolation is adiabatic, because we expect such an approximation to be valid when we can neglect the temporal variation of the smooth component relative to that of matter. Indeed we will see in all our equations below for effective growth exponents that the approximation Eq. (87) becomes exact as ws0w_{s}\rightarrow 0, which corresponds to this limit.

We will apply here this analysis up to third order in perturbation theory, which is what is needed to calculate the one-loop power spectrum. We have chosen for our analysis to follow closely that of [27], who have calculated explicitly the corrections in LCDM and provided phenomenological fits to the relevant redshift dependent functions for standard type models. As we will show, our method of analysis in seeking to relate these results to those in gEdS, leads to a simplification of the formulation and calculations of the results of [27], and other treatments given in the literature (e.g. [11, 28, 20, 32]) Further in so doing it provides greater insight into the reason why these cosmology-dependent corrections are typically so small, and how they depend on cosmological parameters. We compare our numerical results in detail to the fits of [27] in Appendix  D. For completeness we detail also a comparison of our analysis, and numerical results when possible, with that of [32] in Appendix E, and with those of [11] and [20] in Appendix F.

IV.1 Linear growth rate interpolated on gEdS

By combining Eqs.  (11) and (12) at linear order (i.e. neglecting the source terms Sα~S_{\tilde{\alpha}} and Sβ~S_{\tilde{\beta}}) we obtain easily the equation obeyed by the linear growth factor D1(a)D_{1}(a) which can be cast as

d2d2lna2D1a\displaystyle\frac{d^{2}}{d^{2}\ln a^{2}}\frac{D_{1}}{a} +\displaystyle+ (4+dlnHdlna)ddlnaD1a\displaystyle\Big{(}4+\frac{d\ln H}{d\ln a}\Big{)}\frac{d}{d\ln a}\frac{D_{1}}{a} (88)
+\displaystyle+ (3+dlnHdlna32Ωm(a))D1a=0.\displaystyle\Big{(}3+\frac{d\ln H}{d\ln a}-\frac{3}{2}\Omega_{m}(a)\Big{)}\frac{D_{1}}{a}=0.

Writing

D1=exp(lna0lnaα1(a)d(lna0)),D_{1}=\exp\left(\int_{\ln a_{0}}^{\ln a}\alpha_{1}(a)\,d(\ln a_{0})\right), (89)

where a0a_{0} to be the reference time at which D1=1D_{1}=1, and

α1=dlnD1dlna,\alpha_{1}=\frac{d\ln D_{1}}{d\ln a}, (90)

we have, trivially, an interpolation in the sense of Eq. (85) taking αeff(z)=α1\alpha_{eff}(z)=\alpha_{1}.

Let us now consider the accuracy of an adiabatic gEdS interpolation, in the sense defined by Eq. (86) above, i.e., the approximation

α1(z)α10(z)=14[1+1+24Ωm(z)].\alpha_{1}(z)\approx\alpha_{10}(z)=\frac{1}{4}\big{[}-1+\sqrt{1+24\Omega_{m}(z)}\big{]}. (91)

Fig. 2 shows a comparison of the exact growth rate α1\alpha_{1} (obtained by solving Eq. (88) numerically) and α10\alpha_{10}, for models with constant equation of state ws=w0w_{s}=w_{0}, for w0=0.5,1,1.5w_{0}=-0.5,-1,-1.5. The left panel shows the two quantities, and the right panel their ratio. We see that, even for the LCDM case (w0=1w_{0}=-1), the approximation is quite good down to z=0z=0 (when Ωm0.3\Omega_{m}\approx 0.3).

Refer to caption
Refer to caption
Figure 2: Left panel: Exact solution for the growth rate α1\alpha_{1} compared with its “adiabatic gEdS approximation” α10\alpha_{10}, as a function of Ωm\Omega_{m} and for different indicated values of w0w_{0}, parametrizing the constant equation of state of the smooth (dark energy) component. Right panel: the relative difference of the same quantities.

This result is somewhat surprising as one would anticipate that it would hold only for |w0||w_{0}| small compared to unity. To understand what is observed better, we re-express Eq. (88) as an equation for α1\alpha_{1}:

[dα1dlna32wα1]+α12+12α132Ωm=0,\left[\frac{d\alpha_{1}}{d\ln a}-\frac{3}{2}w\alpha_{1}\right]+\alpha_{1}^{2}+\frac{1}{2}\alpha_{1}-\frac{3}{2}\Omega_{m}=0, (92)

where

w=123dlnHdlna=13dlnΩmdlna,w=-1-\frac{2}{3}\frac{d\ln H}{d\ln a}=\frac{1}{3}\frac{d\ln\Omega_{m}}{d\ln a}, (93)

and for the specific case of a smooth (dark energy) component with the constant equation of state, ws=w0w_{s}=w_{0}, we have

w=w0(1Ωm).w=w_{0}(1-\Omega_{m}). (94)

The approximate solution Eq. (91) corresponds to neglecting the first two terms in square brackets i.e., setting w=0w=0 and neglecting the time derivative, and taking the growing mode solution of the remaining quadratic equation. It is thus formally the solution at leading order in a gradient expansion in Ωm(z)\Omega_{m}(z) for which ww (or w0w_{0} for a constant equation of state) is the control parameter. Given that e.g. in LCDM w=0.7w=-0.7 when Ωm=0.3\Omega_{m}=0.3, it is indeed, as we have noted, surprising that the leading w=0w=0 approximation is so good. To see why this is the case more explicitly we define

α1=α10(1+ϵ1(z)).\alpha_{1}=\alpha_{10}(1+\epsilon_{1}(z))\,. (95)

and rewrite Eq. (92), neglecting the single quadratic term in ϵ1\epsilon_{1}, as a linearized equation in ϵ1\epsilon_{1}:

dϵ1dlna+[S(Ωm)w+(12+2α10)]ϵ1=S(Ωm)w,\frac{d\epsilon_{1}}{d\ln a}+\Big{[}-S(\Omega_{m})w+(\frac{1}{2}+2\alpha_{10})\Big{]}\epsilon_{1}=S(\Omega_{m})w, (96)

where

S(Ωm)=321wdlnα10dlna=329Ωmα10(1+4α10),S(\Omega_{m})=\frac{3}{2}-\frac{1}{w}\frac{d\ln\alpha_{10}}{d\ln a}=\frac{3}{2}-\frac{9\Omega_{m}}{\alpha_{10}(1+4\alpha_{10})}, (97)

is the coefficient multiplying ww which gives the source term for ϵ1\epsilon_{1}, the correction to the leading adiabatic gEdS approximation.

Why ϵ1\epsilon_{1} remains small for relatively large ww is just a consequence of the smallness of the source term SS relative to the coefficient of the ϵ1\epsilon_{1} term on the left-hand size of the equation. This is shown explicitly in Fig. 3 which shows the numerical value of S(Ωm)S(\Omega_{m}) plotted against that of (12+2α10)(\frac{1}{2}+2\alpha_{10}).

Refer to caption
Figure 3: The coefficient S(Ωm)S(\Omega_{m}) of the source term in Eq. (96) as a function of (1/2+2α10)(1/2+2\alpha_{10}), and as a function of Ωm\Omega_{m} (upper xx-axis). We see that the ratio of |S(Ωm)|/(1/2+2α10)|S(\Omega_{m})|/(1/2+2\alpha_{10}) remains quite small as dark energy starts to dominate, so that even for |w|1|w|\sim 1 the adiabatic gEdS approximation for the linear growth factor can remain good.

The relative smallness of SS arises from an approximate cancellation of the two terms in Eq. (97), which corresponds to that between the two terms in square brackets in Eq. (92). Indeed at high redshift we have S3295=0.3S\approx\frac{3}{2}-\frac{9}{5}=-0.3.

A corollary of this observed accuracy of the leading gEdS interpolation is that the approximation for α1\alpha_{1} obtained by solving the linear Eq. (96) is an accurate one. Fig. 4 shows a comparison of the value of the growth rate obtained in this way with the exact α1\alpha_{1}. We see that it is indeed accurate well below the sub-percent level even for a LCDM model at z=0z=0.

Refer to caption
Figure 4: Ratio of the growth rate α11\alpha_{11} obtained using the linearized approximation Eq. (96) for ϵ1=(α1/α10)1\epsilon_{1}=(\alpha_{1}/\alpha_{10})-1, compared to the exact solution from Eq. (92).

IV.2 Second-order growth rates interpolated on gEdS models

In treating LCDM models in perturbation theory at second and higher order, a considerable additional complexity arises because the separability of the EdS solution is not valid. In practice it has been shown (see e.g. [14, 28, 20] and references therein) that the associated corrections are very small for standard type cosmological models, and for most purposes can be done using the EdS kernels and the assumption of separability, replacing simply the EdS linear theory growth rate by that of the model. Given the ever-improving precision of cosmological observations, however, even small theoretical errors arising from cosmology dependence are of interest and have been discussed in a number of recent works (see e.g. [46, 47, 48, 49, 20, 26, 30]).

We discuss here the calculations of such corrections in light of our results in gEdS models. More specifically we consider whether the time-dependent functions, additional to D1(a)D_{1}(a), which appear in describing the cosmology dependence of the evolved density field can be interpolated on gEdS models in the sense we have defined at the beginning of this section. In this subsection we consider the density fluctuation at second order, and in the following one at third order. For conciseness, we will not present here the full derivation of the results we use from EPT including cosmological corrections as it has been given in several other works, both for LCDM-type models ([27, 20, 32]) or in a broader class of models (e.g. [28, 17]). As anticipated above we will follow in the main text the treatment of [27], but we also provide in Appendices E-F detailed comparison with several other works [32, 11, 20].

The solution for the density perturbation in a general FLRW cosmology at second order is given by [27] as

δ2(k,a)=D2AA(k)+D2BB(k),\delta_{2}(\textbf{k},a)=D_{2A}A(\textbf{k})+D_{2B}B(\textbf{k}), (98)

where

A(k)=57A^(k)=57d3qα~(q,kq)δ1(q)δ1(kq),\displaystyle A(\textbf{k})=\frac{5}{7}\hat{A}(\textbf{k})=\frac{5}{7}\int d^{3}\textbf{q}\tilde{\alpha}(\textbf{q},\textbf{k}-\textbf{q})\delta_{1}(\textbf{q})\delta_{1}(\textbf{k}-\textbf{q}),
B(k)=27B^(k)=27d3qβ~(q,kq)δ1(q)δ1(kq),\displaystyle B(\textbf{k})=\frac{2}{7}\hat{B}(\textbf{k})=\frac{2}{7}\int d^{3}\textbf{q}\tilde{\beta}(\textbf{q},\textbf{k}-\textbf{q})\delta_{1}(\textbf{q})\delta_{1}(\textbf{k}-\textbf{q}),

and D2AD_{2A} and D2BD_{2B} are time-dependent functions which are solutions to the equations

d2dlna2D2a2+(6+dlnHdlna)ddlnaD2a2\displaystyle\frac{d^{2}}{d\ln a^{2}}\frac{D_{2}}{a^{2}}+\Big{(}6+\frac{d\ln H}{d\ln a}\Big{)}\frac{d}{d\ln a}\frac{D_{2}}{a^{2}} (100)
+[8+2dlnHdlna32Ωm(a)]D2a2\displaystyle+\Big{[}8+2\frac{d\ln H}{d\ln a}-\frac{3}{2}\Omega_{m}(a)\Big{]}\frac{D_{2}}{a^{2}}
=\displaystyle= {75[(dD1da)2+32Ωm(a)(D1a)2] for D2A,72(dD1da)2 for D2B.\displaystyle\begin{cases}\frac{7}{5}\Bigg{[}\Big{(}\frac{dD_{1}}{da}\Big{)}^{2}+\frac{3}{2}\Omega_{m}(a)\Big{(}\frac{D_{1}}{a}\Big{)}^{2}\Bigg{]}\text{ for $D_{2A}$},\\ \frac{7}{2}\Big{(}\frac{dD_{1}}{da}\Big{)}^{2}\text{ for $D_{2B}$}.\end{cases}

These equations have been written in a form adapted to their comparison with the standard EdS model, for which the right-hand side vanishes giving the solutions D2A=D2B=a2D_{2A}=D_{2B}=a^{2}. It is straightforward to verify that in the gEdS cosmology they admit the solutions

D2A=751+4α1+6αa2α,D2B=722α1+6αa2α,D_{2A}=\frac{7}{5}\frac{1+4\alpha}{1+6\alpha}a^{2\alpha},\quad D_{2B}=\frac{7}{2}\frac{2\alpha}{1+6\alpha}a^{2\alpha}, (101)

corresponding exactly to the separable solutions we have derived in the previous section.

In light of these solutions it is natural to define the functions

d2A=57D2AD12,d2B=27D2BD12,d_{2A}=\frac{5}{7}\frac{D_{2A}}{D_{1}^{2}},\quad d_{2B}=\frac{2}{7}\frac{D_{2B}}{D_{1}^{2}}, (102)

in terms of which

δ2(k,a)=d2AD12A^(k)+d2BD12B^(k).\delta_{2}(\textbf{k},a)=d_{2A}D_{1}^{2}\hat{A}(\textbf{k})+d_{2B}D_{1}^{2}\hat{B}(\textbf{k})\,. (103)

When the functions d2Ad_{2A} and d2Bd_{2B} are constant in time, δ2(k,a)\delta_{2}(\textbf{k},a) is a time-independent functional of δ1(k,a)\delta_{1}(\textbf{k},a). Non-separability, on the contrary, corresponds to a time-dependence of the functions d2Ad_{2A} and d2Bd_{2B}. We first note that Eqs. (100) can be written, making use of the definitions of α1\alpha_{1} and ww in the previous section, and using Eq. (92) to eliminate terms in dlnα1/dlnad\ln\alpha_{1}/d\ln a, as

𝒟(2)d2A=α12+32Ωm,𝒟(2)d2B=α12,{\cal{D}}^{(2)}d_{2A}=\alpha_{1}^{2}+\frac{3}{2}\Omega_{m},\quad{\cal{D}}^{(2)}d_{2B}=\alpha_{1}^{2}, (104)

where

𝒟(2)=d2dlna2+[12(13w)+4α1]ddlna+(2α12+32Ωm).{\cal{D}}^{(2)}=\frac{d^{2}}{d\ln a^{2}}+\Big{[}\frac{1}{2}(1-3w)+4\alpha_{1}\Big{]}\frac{d}{d\ln a}+(2\alpha_{1}^{2}+\frac{3}{2}\Omega_{m}). (105)

We see immediately that 𝒟(2)(d2A+d2B1)=0{\cal{D}}^{(2)}(d_{2A}+d_{2B}-1)=0, from which it follows that the evident property

d2A+d2B=1,d_{2A}+d_{2B}=1, (106)

of the usual EdS solution (with d2A=5/7d_{2A}=5/7 and d2B=2/7d_{2B}=2/7), shared by the gEdS cosmologies (with d2A=1+4α1+6αd_{2A}=\frac{1+4\alpha}{1+6\alpha} and d2B=2α1+6αd_{2B}=\frac{2\alpha}{1+6\alpha}), will hold in a general LCDM-like cosmology, if at asymptotically early times the cosmology approaches EdS (or gEdS). There is therefore in this case only one real function (in addition to D1D_{1}) is needed to describe the evolution of the density perturbation at second order, just as in the original treatment of [11]. Given our gEdS solutions (101), it is natural to define, without loss of generality, the time dependence using the function α2(a)\alpha_{2}(a) defined by

d2A(a)=1+4α2(a)1+6α2(a),d2B(a)=1d2A(a).d_{2A}(a)=\frac{1+4\alpha_{2}(a)}{1+6\alpha_{2}(a)},\quad d_{2B}(a)=1-d_{2A}(a)\,. (107)

α2(a)\alpha_{2}(a) is just then the value of α\alpha in the gEdS model with the same instantaneous values of d2Ad_{2A} and d2Bd_{2B}, i.e., it is the effective growth exponent required to interpolate (exactly) these functions on gEdS models. Anticipating the fact that we will consider in particular here the small corrections due to the non-separability of the LCDM model relative to the EdS approximation (with d2A=5/7d_{2A}=5/7 and d2B=2/7d_{2B}=2/7), we also define

d2A=57[1+235γ2],d2B=27[117γ2],d_{2A}=\frac{5}{7}\Big{[}1+\frac{2}{35}\gamma_{2}\Big{]},\quad d_{2B}=\frac{2}{7}\Big{[}1-\frac{1}{7}\gamma_{2}\Big{]}, (108)

where

γ2=1α21+67(1α2),\gamma_{2}=\frac{1-\alpha_{\rm 2}}{1+\frac{6}{7}(1-\alpha_{\rm 2})}\,, (109)

i.e. the coefficients in the definition of the perturbation γ2\gamma_{2} have been chosen so that, in the limit of small γ2\gamma_{2}, we obtain γ2=1α2\gamma_{2}=1-\alpha_{2}.

From Eq. (104) we have that the equation obeyed by γ2\gamma_{2} is

𝒟(2)γ2=212(Ωmα12).{\cal{D}}^{(2)}\gamma_{2}=\frac{21}{2}\Big{(}\Omega_{m}-\alpha_{1}^{2}\Big{)}. (110)

It is instructive to rewrite this equation as

d2γ2dlna2\displaystyle\frac{d^{2}\gamma_{2}}{d\ln a^{2}} +\displaystyle+ [12(13w)+4α1]dγ2dlna\displaystyle\Big{[}\frac{1}{2}(1-3w)+4\alpha_{1}\Big{]}\frac{d\gamma_{2}}{d\ln a} (111)
+\displaystyle+ (2α12+32Ωm)[γ2γ2(0)]=0,\displaystyle\Big{(}2\alpha_{1}^{2}+\frac{3}{2}\Omega_{m}\Big{)}[\gamma_{2}-\gamma_{2}^{(0)}]=0,

where

γ2(0)=21(Ωmα12)4α12+3Ωm.\gamma_{2}^{(0)}=\frac{21(\Omega_{m}-\alpha_{1}^{2})}{4\alpha_{1}^{2}+3\Omega_{m}}\,. (112)

We recover explicitly from this equation the result [38, 39] that the solution is separable if and only if the ratio Ωm/α12\Omega_{m}/\alpha_{1}^{2} is constant. Indeed, as we have noted, separability corresponds to a time-independent solution for γ2\gamma_{2}, which here is admitted if and only if γ2(0)\gamma_{2}^{(0)} is constant in time, and therefore if and only if the ratio Ωm/α12\Omega_{m}/\alpha_{1}^{2} is so. More generally we note that solving in the adiabatic gEdS approximation, i.e. neglecting all terms proportional to ww and all time derivatives, and taking, in the same approximation, α1=α10\alpha_{1}=\alpha_{10}, we then obtain at leading order

γ2=71α101+6α10,\gamma_{2}=7\frac{1-\alpha_{10}}{1+6\alpha_{10}}, (113)

which corresponds to

α2=α10,\alpha_{\rm 2}=\alpha_{10}, (114)

i.e. we recover as expected, as w0w\rightarrow 0, the adiabatic gEdS approximation for d2Ad_{2A} (and d2Bd_{2B}).

Refer to caption
Figure 5: The effective growth rate α2\alpha_{2} of a gEdS model with the same instantaneous value of the functions d2Ad_{2A} (and d2Bd_{2B}), for dark energy models with the constant equation of state parametrized by the indicated values of w0w_{0}. We see that, for very small absolute values of w0w_{0}, α2\alpha_{2} is well approximated by the adiabatic linear growth rate α10\alpha_{10} so that, as expected, d2Ad_{2A} and d2Bd_{2B} are well approximated by an adiabatic interpolation of gEdS models. For larger absolute values of w0w_{0}, α2\alpha_{2} stays much closer to its initial value because of rapid evolution which causes effective damping in its evolution equation.

Fig. 5 shows the numerical solutions obtained for α2\alpha_{2} as a function of Ωm\Omega_{m}, for models with the different indicated values of w0w_{0}. These are obtained by solving Eq. (110) for γ2\gamma_{2}, with initial conditions corresponding to the asymptote to EdS as zz\rightarrow\infty:

γ2(0)=0,dγ2dlna(0)=0.\gamma_{2}(0)=0,\quad\frac{d\gamma_{2}}{d\ln a}(0)=0\,. (115)

In practice, rather than extrapolating to an initial time ai1a_{i}\ll 1 at which Ωm1\Omega_{m}\approx 1, we can avoid the very early time transients by taking directly as the initial condition

γ2(ai)=1α10(ai),dγ2dlna(ai)=dα10dlna(ai),\gamma_{2}(a_{i})=1-\alpha_{10}(a_{i}),\quad\frac{d\gamma_{2}}{d\ln a}(a_{i})=-\frac{d\alpha_{10}}{d\ln a}(a_{i})\,, (116)

since γ2(a)=1α10(a)\gamma_{2}(a)=1-\alpha_{10}(a), as we have noted above, becomes a good approximation to the exact evolution in the limit w1w\ll 1. Indeed this can be seen in Fig. 5, in which this adiabatic approximation α2=α10\alpha_{2}=\alpha_{10} is shown (solid line). We see that, for very small absolute values of w0w_{0}, α2\alpha_{2} is very well approximated by α10\alpha_{10}. In other words as expected, d2Ad_{2A} and d2Bd_{2B} are then well approximated by an adiabatic interpolation of gEdS models. However, unlike what we observed for α1(z)\alpha_{1}(z), this adiabatic approximation fails for increasing values of |w0||w_{0}|. In this case we see that α2\alpha_{2} stays much closer to its initial value. This can be understood simply as a result of the increasing contribution of the dark energy, which increases (through the dependence in ww) an effective damping term in Eq. (110) which slows greatly the “relaxation” to γ2=γ2(0)\gamma_{2}=\gamma_{2}^{(0)} occurring for small |w0||w_{0}|.

IV.3 Third order growth rates

Following again the treatment of [27], we write the density perturbation to third-order as

δ3(k,a)\displaystyle\delta_{3}(\textbf{k},a) =\displaystyle= D3AA(a)CAA(k)+D3AA(a)CAA(k)\displaystyle D_{3AA}(a)C_{AA}(\textbf{k})+D^{\prime}_{3AA}(a)C^{\prime}_{AA}(\textbf{k})
+D3AB(a)CAB(k)+D3AB(a)CAB(k)\displaystyle+D_{3AB}(a)C_{AB}(\textbf{k})+D^{\prime}_{3AB}(a)C^{\prime}_{AB}(\textbf{k})
+D3BA(a)CBA(k)+D3BB(a)CBB(k),\displaystyle+D_{3BA}(a)C_{BA}(\textbf{k})+D_{3BB}(a)C_{BB}(\textbf{k}),

where

CAA(k)\displaystyle C_{AA}(\textbf{k}) =\displaystyle= 718d3qα~(q,kq)δ1(q)A(kq),\displaystyle\frac{7}{18}\int d^{3}\textbf{q}\tilde{\alpha}(\textbf{q},\textbf{k}-\textbf{q})\delta_{1}(\textbf{q})A(\textbf{k}-\textbf{q}),
CAA(k)\displaystyle C_{AA}^{\prime}(\textbf{k}) =\displaystyle= 730d3qα~(q,kq)δ1(kq)A(q),\displaystyle\frac{7}{30}\int d^{3}\textbf{q}\tilde{\alpha}(\textbf{q},\textbf{k}-\textbf{q})\delta_{1}(\textbf{k}-\textbf{q})A(\textbf{q}),
CAB(k)\displaystyle C_{AB}(\textbf{k}) =\displaystyle= 718d3qα~(q,kq)δ1(q)B(kq),\displaystyle\frac{7}{18}\int d^{3}\textbf{q}\tilde{\alpha}(\textbf{q},\textbf{k}-\textbf{q})\delta_{1}(\textbf{q})B(\textbf{k}-\textbf{q}),
CAB(k)\displaystyle C_{AB}^{\prime}(\textbf{k}) =\displaystyle= 79d3qα~(q,kq)δ1(kq)B(q),\displaystyle\frac{7}{9}\int d^{3}\textbf{q}\tilde{\alpha}(\textbf{q},\textbf{k}-\textbf{q})\delta_{1}(\textbf{k}-\textbf{q})B(\textbf{q}),
CBA(k)\displaystyle C_{BA}(\textbf{k}) =\displaystyle= 215d3qβ~(q,kq)δ1(q)A(kq),\displaystyle\frac{2}{15}\int d^{3}\textbf{q}\tilde{\beta}(\textbf{q},\textbf{k}-\textbf{q})\delta_{1}(\textbf{q})A(\textbf{k}-\textbf{q}),
CBB(k)\displaystyle C_{BB}(\textbf{k}) =\displaystyle= 49d3qβ~(q,kq)δ1(q)B(kq).\displaystyle\frac{4}{9}\int d^{3}\textbf{q}\tilde{\beta}(\textbf{q},\textbf{k}-\textbf{q})\delta_{1}(\textbf{q})B(\textbf{k}-\textbf{q}).

and the six functions D3XXD_{3XX} are the solutions of the four differential equations

d2dlna2D3a3+(8+dlnHdlna)ddlnaD3a3\displaystyle\frac{d^{2}}{d\ln a^{2}}\frac{D_{3}}{a^{3}}+\Big{(}8+\frac{d\ln H}{d\ln a}\Big{)}\frac{d}{d\ln a}\frac{D_{3}}{a^{3}} (124)
+[15+3dlnHdlna32Ωm(a)]D3a3\displaystyle+\Big{[}15+3\frac{d\ln H}{d\ln a}-\frac{3}{2}\Omega_{m}(a)\Big{]}\frac{D_{3}}{a^{3}}
=\displaystyle= {187[2dD1da+32Ωm(a)D1a]D2A,Ba2+187adD1daddaD2A,Ba2 for D3AA,3AB15dD1da[addaD2Aa2+2D2Aa275D1adD1da] for D3BA92dD1da[addaD2Ba2+2D2Ba2] for D3BB,\displaystyle\begin{cases}\frac{18}{7}\Big{[}2\frac{dD_{1}}{da}+\frac{3}{2}\Omega_{m}(a)\frac{D_{1}}{a}\Big{]}\frac{D_{2A,B}}{a^{2}}\\ +\frac{18}{7}a\frac{dD_{1}}{da}\frac{d}{da}\frac{D_{2A,B}}{a^{2}}\text{ for $D_{3AA,3AB}$}\\ 15\frac{dD_{1}}{da}\Big{[}a\frac{d}{da}\frac{D_{2A}}{a^{2}}+2\frac{D_{2A}}{a^{2}}-\frac{7}{5}\frac{D_{1}}{a}\frac{dD_{1}}{da}\Big{]}\text{ for $D_{3BA}$}\\ \frac{9}{2}\frac{dD_{1}}{da}\Big{[}a\frac{d}{da}\frac{D_{2B}}{a^{2}}+2\frac{D_{2B}}{a^{2}}\Big{]}\text{ for $D_{3BB}$},\end{cases}

and the two remaining functions are determined by the relations

518D3AA+29D3AB=12D13,\displaystyle\frac{5}{18}D_{3AA}+\frac{2}{9}D_{3AB}^{\prime}=\frac{1}{2}D_{1}^{3},
16D3AA+19D3AB+221D3BA+863D3BB=12D13.\displaystyle\frac{1}{6}D_{3AA}^{\prime}+\frac{1}{9}D_{3AB}+\frac{2}{21}D_{3BA}+\frac{8}{63}D_{3BB}=\frac{1}{2}D_{1}^{3}.

Analogously to the previous section we now define

d3AA\displaystyle d_{3AA} =\displaystyle= 59D3AAD13,d3AA=13D3AAD13,\displaystyle\frac{5}{9}\frac{D_{3AA}}{D_{1}^{3}},\quad d_{3AA}^{\prime}=\frac{1}{3}\frac{D_{3AA}^{\prime}}{D_{1}^{3}},
d3AB\displaystyle d_{3AB} =\displaystyle= 29D3ABD13,d3AB=49D3ABD13,\displaystyle\frac{2}{9}\frac{D_{3AB}}{D_{1}^{3}},\quad d_{3AB}^{\prime}=\frac{4}{9}\frac{D_{3AB}^{\prime}}{D_{1}^{3}},
d3BA\displaystyle\quad d_{3BA} =\displaystyle= 221D3BAD13,d3BB=863D3BBD13.\displaystyle\frac{2}{21}\frac{D_{3BA}}{D_{1}^{3}},\quad d_{3BB}=\frac{8}{63}\frac{D_{3BB}}{D_{1}^{3}}. (126)

For the case of the gEdS cosmologies, it is simple to verify using the kernels F2F_{2}, G2G_{2} and F3F_{3} derived in Section II that we obtain

d3AA\displaystyle d_{3AA} =\displaystyle= d3d2=1+4α1+8α,\displaystyle d_{3}d_{2}=\frac{1+4\alpha}{1+8\alpha},
d3AA\displaystyle d^{\prime}_{3AA} =\displaystyle= d3d~2=1+2α1+8α,\displaystyle d_{3}\tilde{d}_{2}=\frac{1+2\alpha}{1+8\alpha},
d3AB\displaystyle d_{3AB} =\displaystyle= d3(1d2)=2α1+8α,\displaystyle d_{3}(1-d_{2})=\frac{2\alpha}{1+8\alpha},
d3AB\displaystyle d^{\prime}_{3AB} =\displaystyle= d3(1d~2)=4α1+8α,\displaystyle d_{3}(1-\tilde{d}_{2})=\frac{4\alpha}{1+8\alpha},
d3BA\displaystyle d_{3BA} =\displaystyle= (1d3)d~2=2α(1+2α)(1+6α)(1+8α),\displaystyle(1-d_{3})\tilde{d}_{2}=\frac{2\alpha(1+2\alpha)}{(1+6\alpha)(1+8\alpha)},
d3BB\displaystyle d_{3BB} =\displaystyle= (1d3)(1d~2)=8α2(1+6α)(1+8α).\displaystyle(1-d_{3})(1-\tilde{d}_{2})=\frac{8\alpha^{2}}{(1+6\alpha)(1+8\alpha)}.

It is straightforward to show that using Eqs. (124), and d2B=1d2Ad_{2B}=1-d_{2A}, we obtain

𝒟(3)d3AA\displaystyle{\cal{D}}^{(3)}d_{3AA} =\displaystyle= (4α12+3Ωm)d2A+2α1dd2Adlna,\displaystyle(4\alpha_{1}^{2}+3\Omega_{m})d_{2A}+2\alpha_{1}\frac{d\,d_{2A}}{d\ln a},
𝒟(3)d3AB\displaystyle{\cal{D}}^{(3)}d_{3AB} =\displaystyle= (4α12+3Ωm)(4α12+3Ωm)d2A\displaystyle(4\alpha_{1}^{2}+3\Omega_{m})-(4\alpha_{1}^{2}+3\Omega_{m})d_{2A}
2α1dd2Adlna,\displaystyle-2\alpha_{1}\frac{d\,d_{2A}}{d\ln a},
𝒟(3)d3BA\displaystyle{\cal{D}}^{(3)}d_{3BA} =\displaystyle= 2α12+4α12d2A+2α1dd2Adlna,\displaystyle-2\alpha_{1}^{2}+4\alpha_{1}^{2}d_{2A}+2\alpha_{1}\frac{d\,d_{2A}}{d\ln a},
𝒟(3)d3BB\displaystyle{\cal{D}}^{(3)}d_{3BB} =\displaystyle= 4α124α12d2A2α1dd2Adlna,\displaystyle 4\alpha_{1}^{2}-4\alpha_{1}^{2}d_{2A}-2\alpha_{1}\frac{d\,d_{2A}}{d\ln a}, (128)

where

𝒟(3)=d2dlna2+[12(13w)+6α1]ddlna+3(2α12+Ωm).{\cal{D}}^{(3)}=\frac{d^{2}}{d\ln a^{2}}+\Big{[}\frac{1}{2}(1-3w)+6\alpha_{1}\Big{]}\frac{d}{d\ln a}+3(2\alpha_{1}^{2}+\Omega_{m})\,. (129)

The two constraints Eqs. (IV.3) may be written as

d3AA+d3AB=1,\displaystyle d_{3AA}+d_{3AB}^{\prime}=1, (130)
d3AB+d3AA+2(d3BA+d3BB)=1.\displaystyle d_{3AB}+d_{3AA}^{\prime}+2(d_{3BA}+d_{3BB})=1. (131)

Setting all derivatives to zero in Eq. (IV.3), we recover, as required, the expression for the gEdS cosmologies Eq. (IV.3) above.

Taking the sum of Eqs. (IV.3) we note that we obtain also

𝒟(3)(d3AA+d3AB+d3BA+d3BB1)=0,{\cal{D}}^{(3)}(d_{3AA}+d_{3AB}+d_{3BA}+d_{3BB}-1)=0, (132)

and therefore, given that this quantity is zero in an EdS (and gEdS) cosmology, we obtain the additional constraint

d3AA+d3AB+d3BA+d3BB=1,\displaystyle d_{3AA}+d_{3AB}+d_{3BA}+d_{3BB}=1, (133)

when we assume that the cosmology is asymptotically EdS (or gEdS). Further writing Eq. (104) for d2Ad_{2A} as

𝒟(3)d2A=α12+32Ωm+(4α12+32Ωm)d2A+2α1dd2Adlna,{\cal{D}}^{(3)}d_{2A}=\alpha_{1}^{2}+\frac{3}{2}\Omega_{m}+(4\alpha_{1}^{2}+\frac{3}{2}\Omega_{m})d_{2A}+2\alpha_{1}\frac{d\,d_{2A}}{d\ln a}, (134)

we see also that

𝒟(3)(2d2Ad3AA+d3BB1)=0,{\cal{D}}^{(3)}(2d_{2A}-d_{3AA}+d_{3BB}-1)=0, (135)

so that we can infer (again assuming the cosmology to be asymptotically EdS or gEdS) the additional relation

2d2Ad3AA+d3BB=1.2d_{2A}-d_{3AA}+d_{3BB}=1. (136)

The full set of constraints now reads

d3AA+d3AB=1,\displaystyle d_{3AA}+d_{3AB}^{\prime}=1, (137)
d3AB+d3AA+d3BA+d3BB=1,\displaystyle d_{3AB}^{\prime}+d_{3AA}^{\prime}+d_{3BA}+d_{3BB}=1, (138)
d3AB+d3AA+d3BA+d3BB=1,\displaystyle d_{3AB}+d_{3AA}+d_{3BA}+d_{3BB}=1, (139)
2d2Ad3AA+d3BB=1.\displaystyle 2d_{2A}-d_{3AA}+d_{3BB}=1. (140)

We can thus conclude that the perturbation at the third order can therefore be described fully by only just two independent functions in addition to d2Ad_{2A}. This result agrees again, as we saw at second order, with the original analysis of [11] for the asymptotic EdS case. The exact mapping between our functions and those of [11] is given in Appendix F. Our analysis above shows that this same result (and either our parametrization or that of [11]) generalizes to the broader class of cosmologies which are asymptotically gEdS.

We choose now (arbitrarily) to use, in addition to d2Ad_{2A}, the two functions d3AAd_{3AA} and d3ABd_{3AB} to describe the cosmology-dependent corrections to the third-order perturbation To describe them in terms of interpolation on gEdS models we thus introduce, following the treatment of the second order perturbation, effective growth exponents as follows:

d3AA(a)\displaystyle d_{3AA}(a) =\displaystyle= 1+4α3AA(a)1+8α3AA(a)=59[1+445γ3AA(a)],\displaystyle\frac{1+4\alpha_{3AA}(a)}{1+8\alpha_{3AA}(a)}=\frac{5}{9}\Big{[}1+\frac{4}{45}\gamma_{3AA}(a)\Big{]},
d3AB(a)\displaystyle d_{3AB}(a) =\displaystyle= 2α3AB(a)1+8α3AB(a)=29[119γ3AB(a)],\displaystyle\frac{2\alpha_{3AB}(a)}{1+8\alpha_{3AB}(a)}=\frac{2}{9}\Big{[}1-\frac{1}{9}\gamma_{3AB}(a)\Big{]},

where, as for the second order case, the numerical coefficients have been chosen so that we will have, in the limit of small γ3XX\gamma_{3XX}, that α3XX1γ3XX\alpha_{3XX}\approx 1-\gamma_{3XX}.

The evolution of the γ3AA\gamma_{3AA} and γ3AB\gamma_{3AB} are now given by the equations

𝒟(3)γ3AA\displaystyle{\cal{D}}^{(3)}\gamma_{3AA} =\displaystyle= 13514(Ωmα12)\displaystyle\frac{135}{14}\Big{(}\Omega_{m}-\alpha_{1}^{2}\Big{)} (142)
+8198[(4α12+3Ωm)γ2+2α1dγ2dlna],\displaystyle+\frac{81}{98}\Big{[}(4\alpha_{1}^{2}+3\Omega_{m})\gamma_{2}+2\alpha_{1}\frac{d\,\gamma_{2}}{d\ln a}\Big{]},\
𝒟(3)γ3AB\displaystyle{\cal{D}}^{(3)}\gamma_{3AB} =\displaystyle= 547(Ωmα12)\displaystyle-\frac{54}{7}\Big{(}\Omega_{m}-\alpha_{1}^{2}\Big{)}
+8149[(4α12+3Ωm)γ2+2α1dγ2dlna],\displaystyle+\frac{81}{49}\Big{[}(4\alpha_{1}^{2}+3\Omega_{m})\gamma_{2}+2\alpha_{1}\frac{d\,\gamma_{2}}{d\ln a}\Big{]},\

which can also be written as

d2γ3XXdlna2+[12(13w)+6α1]dγ3XXdlna+3[2α12+Ωm]×(γ3XXγ3XX(0))=c3XX[(4α12+3Ωm)×(γ2γ2(0))+2α1dγ2dlna],\begin{split}\frac{d^{2}\gamma_{3XX}}{d\ln a^{2}}+&\Big{[}\frac{1}{2}(1-3w)+6\alpha_{1}\Big{]}\frac{d\gamma_{3XX}}{d\ln a}+3\Big{[}2\alpha_{1}^{2}+\Omega_{m}\Big{]}\\ &\times(\gamma_{3XX}-\gamma_{3XX}^{(0)})=c_{3XX}\Big{[}(4\alpha_{1}^{2}+3\Omega_{m})\\ &\times(\gamma_{2}-\gamma_{2}^{(0)})+2\alpha_{1}\frac{d\,\gamma_{2}}{d\ln a}\Big{]},\end{split} (144)

where c3AA=8198,c3AB=8149c_{3AA}=\frac{81}{98},c_{3AB}=\frac{81}{49} and

γ3AA(0)=γ3AB(0)=9(Ωmα12)2α12+Ωm.\gamma_{3AA}^{(0)}=\gamma_{3AB}^{(0)}=\frac{9(\Omega_{m}-\alpha_{1}^{2})}{2\alpha_{1}^{2}+\Omega_{m}}. (145)

For purposes of comparison it is useful to rewrite Eq. (111) for γ2\gamma_{2} as

𝒟(3)γ2=212(Ωmα12)+[(4α12+32Ωm)γ2+2α1dγ2dlna],\begin{split}{\cal{D}}^{(3)}\gamma_{2}&=\frac{21}{2}\Big{(}\Omega_{m}-\alpha_{1}^{2}\Big{)}+\Big{[}(4\alpha_{1}^{2}+\frac{3}{2}\Omega_{m})\gamma_{2}+2\alpha_{1}\frac{d\,\gamma_{2}}{d\ln a}\Big{]},\end{split} (146)

and

d2γ2dlna2+[12(13w)+6α1]dγ2dlna+3[2α12+Ωm]×(γ2γ2(0))=[(4α12+32Ωm)(γ2γ2(0))+2α1dγ2dlna].\begin{split}\frac{d^{2}\gamma_{2}}{d\ln a^{2}}+&\Big{[}\frac{1}{2}(1-3w)+6\alpha_{1}\Big{]}\frac{d\gamma_{2}}{d\ln a}+3\Big{[}2\alpha_{1}^{2}+\Omega_{m}\Big{]}\\ &\times(\gamma_{2}-\gamma_{2}^{(0)})=\Big{[}(4\alpha_{1}^{2}+\frac{3}{2}\Omega_{m})(\gamma_{2}-\gamma_{2}^{(0)})\\ &+2\alpha_{1}\frac{d\,\gamma_{2}}{d\ln a}\Big{]}.\end{split} (147)

Solving Eqs. (146) and (142) in the adiabatic gEdS approximation, i.e. neglecting all time derivatives in these equations, we can see that we obtain the solution Eq. (113) for γ2\gamma_{2} and now also

γ3AA=γ3AB=91α101+8α1,0,\gamma_{3AA}=\gamma_{3AB}=9\frac{1-\alpha_{10}}{1+8\alpha_{1,0}}, (148)

which corresponds, as anticipated, to the leading adiabatic solutions

α2(a)=α3AA(a)=α3AB(a)=α10(a).\alpha_{2}(a)=\alpha_{3AA}(a)=\alpha_{3AB}(a)=\alpha_{10}(a). (149)
Refer to caption
Figure 6: Evolution of the effective growth rate α2\alpha_{2}, α3AA\alpha_{3AA} and α3AB\alpha_{3AB} which describes in general the cosmology-dependent corrections to the separable EdS approximation at third order in perturbation theory. The different sets of curves correspond to the different indicated values of w0w_{0}. Also shown is the adiabatic linear growth rate α10\alpha_{10}. We see that for the smallest values the adiabatic approximation is good, but that it degrades increasingly as ww increases.
Refer to caption
Figure 7: Same as in the previous figure but for models with larger absolute values of w0w_{0}. We see that for these values of w0w_{0} becomes larger, one of the three exponents has quite a different value from the α2\alpha_{2}, so an interpolation of the functions describing deviations from separability on a single gEdS the model with α=α2\alpha=\alpha_{2} is not necessarily a valid approximation in general, although it turns out to be for the calculation of the one-loop PS.

Fig. 6 and Fig. 7 show the results for the three effective growth exponents α2\alpha_{2} (as in Fig. 5), α3AA\alpha_{3AA} and α3AB\alpha_{3AB} as a function of Ωm\Omega_{m} obtained by numerically integrating these equations for different dark energy models with the fixed equation of state with the same initial conditions as for γ2\gamma_{2} (corresponding to EdS at high redshift). For the smallest value of w0w_{0} we see, as expected, all three exponents well approximated by the adiabatic solution Eq. (149). Then as |w0||w_{0}| increases we see that not only do the exponents differ from this solution but they start differing, albeit to a lesser extent, from one another. More specifically we note that α2α3AA\alpha_{2}\approx\alpha_{3AA}, while the third exponent differs. The origin of these behaviours can be seen easily in the coefficients of the equations above: those in the equations for γ2\gamma_{2} and α3AA\alpha_{3AA} are almost identical, while those in the equation α3AB\alpha_{3AB} differ not only in magnitude (by about a factor of 22) but also in sign. In Appendix D we report a detailed comparison of our numerical results with those of [27], and adapt the fitting functions provided by it to infer corresponding fits for the two functions γ3AA\gamma_{3AA} and γ3AB\gamma_{3AB}.

IV.4 The PS at one loop

We now derive our simplified expressions for the PS at one loop in the class of FLRW cosmologies specified in Section II.1 (clustering matter plus a generic dark energy fluid component), making use of the reduction of the number of independent functions and expressing these in terms of the effective growth exponents we have defined above. We also consider finally to what accuracy this exact result can be approximated by a direct interpolation of the PS in gEdS models (i.e. by just replacing α\alpha in the analytic gEdS kernels by a single effective growth rate). In our analysis we start from the expressions we need (for each of the ‘22’ and ‘13’ contributions) in a form easily comparable (up to notation) to the expressions given in [27]. This facilitates the explanation of the simplifications we obtain.

IV.4.1 ‘22’ term

Starting from Eq. (98), we obtain the full time-dependent P22P_{22} directly as

P22\displaystyle P_{22} =\displaystyle= d2A2[M0+M1+M2]+2d2Ad2B[M0+12M1]\displaystyle d_{2A}^{2}[M_{0}+M_{1}+M_{2}]+2d_{2A}d_{2B}[M_{0}+\frac{1}{2}M_{1}] (150)
+d2BB2M0\displaystyle+d_{2BB}^{2}M_{0}
=\displaystyle= (d2A+d2B)2M0+d2A(d2A+d2B)M1\displaystyle(d_{2A}+d_{2B})^{2}M_{0}+d_{2A}(d_{2A}+d_{2B})M_{1}
+d2A2M2\displaystyle+d_{2A}^{2}M_{2}

where the integrals MiM_{i} are those defined above in Eq. (77). The first expression is, up to simple notational differences, identical to that given by [27].222An equivalent expression is given in [28], also in terms of two redshift dependent functions and three integrals which are linear combinations of M0M_{0}, M1M_{1} and M2M_{2}. Making use of the additional relation Eq. (106) we have derived above, this simplifies to

P22=M0+d2AM1+d2A2M2,P_{22}=M_{0}+d_{2A}M_{1}+d_{2A}^{2}M_{2}, (151)

i.e. to a form evidently equivalent to that in the gEdS model Eq. (75) when we parametrize d2Ad_{2A} by α2\alpha_{2} or γ2\gamma_{2} as defined in the previous section. Defining now

ΔP22=P22P22EdS,\Delta P_{22}=P_{22}-P_{22}^{EdS}, (152)

where P22EdSP_{22}^{EdS} is the result in the separable EdS approximation

P22EdS=M0+57M1+2549M2,P_{22}^{EdS}=M_{0}+\frac{5}{7}M_{1}+\frac{25}{49}M_{2}, (153)

the cosmology dependent correction ΔP22\Delta P_{22} is expressed explicitly as a function of γ2(a)\gamma_{2}(a) by

ΔP22=249γ2(M1+107M2)+(249γ2)2M2.\Delta P_{22}=\frac{2}{49}\gamma_{2}(M_{1}+\frac{10}{7}M_{2})+(\frac{2}{49}\gamma_{2})^{2}M_{2}. (154)

Recalling our discussion below we see that this correction is thus given in terms just of integrals that are explicitly infra-red safe, which is not the case if we do not make use of the relation Eq. (106).

Table 2: A summary of the calculation of the three “effective growth rates” parametrizing the second and third-order density kernels using the different approximations we have discussed and also the exact calculation.
Approximation α2\alpha_{2} α3AA\alpha_{3AA} α3AB\alpha_{3AB} Precision
EdS 11 11 11 0.5%\lesssim 0.5\% for PS at z=0z=0 (LCDM)
adiabatic gEdS α1=dlnD1dlna\alpha_{1}=\frac{d\ln D_{1}}{d\ln a}, Eq. (92) α1\alpha_{1} α1\alpha_{1} valid for small |w0||w_{0}| (cf. Figs. 5, 6, and 7)
gEdS α2\alpha_{2}, Eq. (111) α2\alpha_{2} α2\alpha_{2} 0.1%\lesssim 0.1\% for PS at z=0z=0 (LCDM)
exact α2\alpha_{2}, Eq. (111) α3AA\alpha_{3AA} , Eq.(144) α3AB\alpha_{3AB}, Eq. (144)

IV.4.2 ‘13’ term

Starting from Eq. (IV.3), the expression for P13P_{13} can be obtained directly as

2P13\displaystyle 2P_{13} =\displaystyle= d3AA(N0+N3)+d3AA(N1N3)\displaystyle d_{3AA}(N_{0}+N_{3})+d_{3AA}^{\prime}(N_{1}-N_{3}) (155)
+d3AB(N0+N3)d3ABN3+d3BA(N0+N2)\displaystyle+d_{3AB}(N_{0}+N_{3})-d_{3AB}^{\prime}N_{3}+d_{3BA}(N_{0}+N_{2})
+d3BBN0\displaystyle+d_{3BB}N_{0}
=\displaystyle= (d3AA+d3AB+d3BA+d3BB)N0\displaystyle(d_{3AA}+d_{3AB}+d_{3BA}+d_{3BB})N_{0}
+d3AAN1+d3BAN2+(d3AAd3AA\displaystyle+d_{3AA}^{\prime}N_{1}+d_{3BA}N_{2}+(d_{3AA}-d_{3AA}^{\prime}
+d3ABd3AB)N3,\displaystyle+d_{3AB}-d_{3AB}^{\prime})N_{3},

where the integrals N0,N1,N2N_{0},N_{1},N_{2} are those defined by Eq. (81) and Eq. (82). N3N_{3} is an additional integral defined also by Eq. (81) but with n3=43r2n_{3}=-\frac{4}{3}r^{2}, and which, it is simple to verify, is infra-red safe. The first expression has been written to be easily compared to that given by [27] in terms of the six functions d3XXd_{3XX} and six integrals.333The equivalent expression in [28] is given in terms of six redshift dependent functions and five integrals. Making use now of the relations Eqs. (137)-(140) obtained from the EdS (or gEdS) boundary condition, we find that the coefficient of N3N_{3} vanishes, and further that the expression can be simplified to be a function of only of d2Ad_{2A} and only one of the initial six functions d3XXd_{3XX}:

2P13=N0N1+2d2AN1+d3BA(N2N1),2P_{13}=N_{0}-N_{1}+2d_{2A}N_{1}+d_{3BA}(N_{2}-N_{1}), (156)

Thus, as for P22P_{22}, we obtain an expression involving only the same three integrals as in the gEdS one-loop PS, and as in that case, the cosmology-dependent terms involve only the two infra-red safe integrals N1N_{1} and N2N_{2}. We note that this latter property is obtained using the relation Eq. (139). Conversely, without employing this relation, which we have shown here follows from the condition that the cosmology asymptotes to EdS (or gEdS) at early times, it is not possible to recover explicitly the infra-red safety of the cosmological correction.

The cosmological correction in P13P_{13} can thus be taken to depend in general only on γ2\gamma_{2}, as defined in Eq. (108), and on just one additional function. It is convenient then to introduce the function γ3BA=γ3\gamma_{3BA}=\gamma_{3}, defined, following the same treatment as in Section IV, as

d3BA=2α3(1+2α3)(1+6α3)(1+8α3)=221[1+563γ3],d_{3BA}=\frac{2\alpha_{3}(1+2\alpha_{3})}{(1+6\alpha_{3})(1+8\alpha_{3})}=\frac{2}{21}\Big{[}1+\frac{5}{63}\gamma_{3}\Big{]}, (157)

so that α31γ3\alpha_{3}\approx 1-\gamma_{3} for γ31\gamma_{3}\ll 1. γ3\gamma_{3} is given in terms of γ2\gamma_{2} and the two functions γ3AA\gamma_{3AA} and γ3BA\gamma_{3BA} analysed in the Section IV, by

15γ3=162γ2196γ3AA+49γ3AB.\displaystyle 15\gamma_{3}=162\gamma_{2}-196\gamma_{3AA}+49\gamma_{3AB}. (158)

It is thus a solution to the equation

𝒟(3)γ3=1895(Ωmα12)+18935[4α12γ2+2α1dγ2dlna],{\cal{D}}^{(3)}\gamma_{3}=-\frac{189}{5}\Big{(}\Omega_{m}-\alpha_{1}^{2}\Big{)}+\frac{189}{35}\Big{[}4\alpha_{1}^{2}\gamma_{2}+2\alpha_{1}\frac{d\,\gamma_{2}}{d\ln a}\Big{]}, (159)

with appropriate boundary conditions. Defining

2ΔP13=2P132P13EdS,2\Delta P_{13}=2P_{13}-2P_{13}^{EdS}, (160)

where

2P13EdS=N0+13N1+221N2,2P_{13}^{EdS}=N_{0}+\frac{1}{3}N_{1}+\frac{2}{21}N_{2}, (161)

is the usual result in the separable EdS approximation, the cosmological correction is given as a function of γ2\gamma_{2} and γ3\gamma_{3} as

2ΔP13=449[γ2N1+554γ3(N2N1)].2\Delta P_{13}=\frac{4}{49}\big{[}\gamma_{2}N_{1}+\frac{5}{54}\gamma_{3}(N_{2}-N_{1})\big{]}. (162)

IV.4.3 The cosmological correction to the PS at one loop

Combining Eqs. (154) and (162), we get the final expression for the full cosmological (i.e. non-EdS) correction to the one-loop PS as

ΔP1loop\displaystyle\Delta P_{1-loop} =\displaystyle= 249γ2(M1+107M2)+(249γ2)2M2\displaystyle\frac{2}{49}\gamma_{2}(M_{1}+\frac{10}{7}M_{2})+(\frac{2}{49}\gamma_{2})^{2}M_{2} (163)
+449[γ2N1+554γ3(N2N1)].\displaystyle+\frac{4}{49}\big{[}\gamma_{2}N_{1}+\frac{5}{54}\gamma_{3}(N_{2}-N_{1})\big{]}.
Refer to caption
Refer to caption
Figure 8: Left panel: The one loop cosmological (non-EdS) correction ΔP1loop\Delta P_{1-loop} as a fraction of P1loopP_{1-loop}, the full one-loop PS in EdS, at z=0, for w0=1w_{0}=-1 and w0=0.5w_{0}=-0.5 (and a typical standard like cosmological PS, see text) calculated using the exact expression Eq. (163), and also using the exact gEdS kernels with α\alpha replaced by the single effective growth exponent α2(z)\alpha_{2}(z). Right panel: ratio of the gEdS approximation to the exact result.

The left panel in Fig. 8 shows ΔP1loop\Delta P_{1-loop} relative to the full one loop PS in EdS, at z=0z=0, in a standard cosmological model (specifically, with the cosmological parameters of [50] and calculated using the approximation to the transfer function of [51]). The model labeled by w0=0.5w_{0}=-0.5 is identical but for this parameter. As previously documented in the literature (e.g. [27]) the corrections are at the sub percent level in the former case but increase as |w0||w_{0}| decreases. The reason for this behaviour is very evident in light of our analysis: a smaller |w||w| is closer to the adiabatic limit, which corresponds to an effective α\alpha closer to the instantaneous linear growth exponent α1=dlogD1/dloga\alpha_{1}=d\log D_{1}/d\log a (i.e. to a larger value of |γ2||\gamma_{2}| and |γ3||\gamma_{3}|.) Also shown are the approximation to ΔP1loop\Delta P_{1-loop} obtained by directly interpolating with a gEdS model i.e., by replacing α\alpha by α2(z)\alpha_{2}(z) in Eqs. (75) and (76) (and subtracting the EdS result from their sum). The right panel of Fig. 8 shows the ratio of this approximation to the exact result. As might be anticipated from the expression Eq. (163) — which depends on the function γ3\gamma_{3} only in one term with a small pre-factor — we find that interpolation provides a good approximation (within about 25%25\%). In practice it is not significantly more difficult to calculate the exact result, but the result shows that for understanding the origin of the smallness of the calculated correction, the approximation to gEdS is very instructive. We underline, however, that the adiabatic approximation α2α10\alpha_{2}\approx\alpha_{10} is not valid, so this is not an adiabatic interpolation in the sense we have defined. The effective growth exponent α2\alpha_{2} can be roughly thought of a time-averaged behaviour of the exact linear growth exponent α1\alpha_{1}, which is itself a good approximation to α2\alpha_{2} for small |w||w|.

To summarize we present in Table 2 a synthesis of the different approximations for the calculation of density perturbations up to third order, in cosmological models with a self-gravitating matter component and a smooth time-dependent component (given through an effective equation of state p=w(a)ρp=w(a)\rho). To parametrize the temporal evolution of the kernels we have defined the three functions α2\alpha_{2}, α3AA\alpha_{3AA} and α3AB\alpha_{3AB} (and also equivalent functions γ2\gamma_{2}, γ3AA\gamma_{3AA} and γ3AB\gamma_{3AB}). The table indicates how each of these three functions is calculated in each of the three approximations we have discussed for the exact result.

V Discussion and conclusions

We have derived in detail the kernels for both standard EPT and LPT in a family of EdS models, characterized by their constant linear growth rate α\alpha. While such results for this family of models are implicit in previous treatments of perturbation theory in the FLRW cosmologies (e.g. [27]), or in a broader class of cosmologies (e.g. [28, 32]), the simple analytical results in this case which generalize those of the standard EdS model have not (to our knowledge) been discussed in the literature previously. For the one-loop power spectrum in EFT we have used these kernels to show that the expected infra-red convergence properties of the corrections relative to a standard EdS model are obtained, without the need for cancellation between the two contributing integrals. Further we have noted that the coefficients of ultra-violet divergent terms have a strong dependence on α\alpha, and the leading divergence can even change sign (and is zero at a certain value).

In the second part of the paper we have shown these analytical results for the kernels in these models can be exploited to obtain a simplified formulation of the calculation of the cosmology-dependent corrections to the usual separable EdS approximation in any standard type non-EdS cosmology, and in particular in LCDM-like cosmologies. We do so by parametrizing the relevant time-dependent functions characterizing this cosmology dependence as effective growth rates in gEdS models. Assuming only that the cosmology at asymptotically early times is gEdS, second-order corrections are parametrized in terms of one such exponent, and third-order corrections in terms of it and two more such exponents. Thus, for example, the bispectrum calculated at leading non-trivial order in perturbation theory, depending on the second-order perturbation, is, at any time, exactly equal to that in a gEdS model. For the power spectrum we show the results on the infra-red convergence behaviour of the gEdS model generalize, and we derive explicit expressions for the corrections to the EdS model expressed in terms of infra-red safe integrals. This expression turns out to depend on only one of the two effective exponents needed at the third order. Nevertheless the PS is in fact given to a relatively good approximation (up to 25%\sim 25\%) by that obtained with the analytic gEdS kernels on replacing the fixed growth rate α\alpha of gEdS by the single effective exponent α2(z)\alpha_{2}(z) defining the exact evolution at second order. Our exact expression is much simplified compared to previous (equivalent) expressions derived in [27, 32] which involve six or eight redshift-dependent functions and do not explicitly recover the infra-red convergent properties as we have done.

Our formulation of the cosmological corrections in this way makes it simple to understand why they are so very small in LCDM-like models (typically <1%<1\% even at z=0z=0). The effective growth exponents we introduce to map the time-dependent kernels to those of the gEdS models obey equations in which one can clearly identify an adiabatic limit, of a dark energy component scaling like matter (i.e. w=0w=0) in which these exponents converge to the instantaneous logarithmic linear growth rate. The more rapid evolution of dark energy at low redshift acts like a damping of this evolution, leading to the effective exponent much closer to its initial EdS value. Combined with the very weak α\alpha dependence shown by the analytical expression for the gEdS kernels, the corrections turn out to be as small as they are. Only for models accelerating much faster than LCDMLCDM at z=0z=0 can such corrections be comparable to the EdS approximated loop corrections, and even in the asymptotic limit that Ωm0\Omega_{m}\rightarrow 0 they remain finite and at most of the order unity.

In a forthcoming paper, we will exploit the results for the kernels in gEdS to obtain exact analytical solutions for the PS in what we will call “generalized scale-free models” i.e. for a PS which is a simple power law and the cosmology is gEdS [34, 33]. These models share the property of self-similarity of usual scale-free models (with a standard EdS expansion law). This makes them ideal for numerical testing for the accuracy of the analytical predictions, as the property of self-similarity allows the extraction of highly accurate results from simulations (see [35, 36]). In further work we plan also to explore the ultra-violet divergences in perturbation theory in these models, and to examine their regularization by methods such as effective field theory. This may, we anticipate, provide insights into the application of such techniques in standard cosmologies. It would also be interesting to extend the analysis we have presented here to determine analogous formulations (in terms of effective growth rate functions) of analyses previously reported in the literature for cosmological corrections, both for other statistics (two-point velocity statistics, pairwise velocities, trispectrum etc.) at one-loop and two-loop, and also for a broader range of cosmological models. In particular it would be interesting to explore such a formulation of perturbation theory with a massive neutrino component including its perturbations (see e.g. [48, 49, 20, 52]). In this case the kernels have also a scale dependence that might be expected to be conveniently formulated in terms of (scale-dependent) effective growth exponents analogous to those we have employed here.

Acknowledgements.
A.P. is supported by the Indonesia Endowment Fund for Education (LPDP). The numerical calculations in this research used the python packages Matplotlib [53], NumPy [54], and SciPy [55].

Appendix A PT kernels at third-order in gEdS

A.1 EPT

Eq. (18) at third order in δ(1)\delta^{(1)} gives

2{a2a232aa\displaystyle\mathcal{H}^{2}\Big{\{}-a^{2}\partial_{a}^{2}-\frac{3}{2}a\partial_{a} +\displaystyle+ 32κ2}δ(3)(k,a)=S(3)β~(k,a)a(aSα~(3)(k,a)),\displaystyle\frac{3}{2\kappa^{2}}\Big{\}}\delta^{(3)}(\textit{{k}},a)=S^{(3)}_{\tilde{\beta}}(\textit{{k}},a)-\mathcal{H}\partial_{a}(aS^{(3)}_{\tilde{\alpha}}(\textit{{k}},a)), (164)

where

Sα~(3)(k,a)=d3q(2π)3α~(q,kq,a)θ(1)(q,a)δ(2)(kq,a)d3q(2π)3α~(q,kq,a)θ(2)(q,a)δ(1)(kq,a),\displaystyle S_{\tilde{\alpha}}^{(3)}(\textit{{k}},a)=-\int\frac{\text{d}^{3}q}{(2\pi)^{3}}\tilde{\alpha}(\textit{{q}},\textit{{k}}-\textit{{q}},a)\theta^{(1)}(\textit{{q}},a)\delta^{(2)}(\textit{{k}}-\textit{{q}},a)-\int\frac{\text{d}^{3}q}{(2\pi)^{3}}\tilde{\alpha}(\textit{{q}},\textit{{k}}-\textit{{q}},a)\theta^{(2)}(\textit{{q}},a)\delta^{(1)}(\textit{{k}}-\textit{{q}},a), (165)
Sβ~(3)(k,a)=d3q(2π)3β~(q,kq,a)θ(1)(q,a)θ(2)(kq,a)d3q(2π)3β~(q,kq,a)θ(2)(q,a)θ(1)(kq,a).\displaystyle S_{\tilde{\beta}}^{(3)}(\textit{{k}},a)=-\int\frac{\text{d}^{3}q}{(2\pi)^{3}}\tilde{\beta}(\textit{{q}},\textit{{k}}-\textit{{q}},a)\theta^{(1)}(\textit{{q}},a)\theta^{(2)}(\textit{{k}}-\textit{{q}},a)-\int\frac{\text{d}^{3}q}{(2\pi)^{3}}\tilde{\beta}(\textit{{q}},\textit{{k}}-\textit{{q}},a)\theta^{(2)}(\textit{{q}},a)\theta^{(1)}(\textit{{k}}-\textit{{q}},a). (166)

We can write Eq. (165) as

Sα~(3)(k,a)=fD3(a)(A1+A2),\begin{split}S^{(3)}_{\tilde{\alpha}}(\textit{{k}},a)=\mathcal{H}fD^{3}(a)(A_{1}+A_{2}),\end{split} (167)

where

A1\displaystyle A_{1} =\displaystyle= d3q(2π)3d3q1(2π)3d3q2(2π)3δ(1)(q)δ(1)(q1)δ(1)(q2)α~(q,kq)F2(q1,q2)(2π)3δ(D)(kqq1q2),\displaystyle\int\frac{\text{d}^{3}q}{(2\pi)^{3}}\frac{\text{d}^{3}q_{1}}{(2\pi)^{3}}\frac{\text{d}^{3}q_{2}}{(2\pi)^{3}}\delta^{(1)}(\textit{{q}})\delta^{(1)}(\textit{{q}}_{1})\delta^{(1)}(\textit{{q}}_{2})\tilde{\alpha}(\textit{{q}},\textit{{k}}-\textit{{q}})\textit{F}_{2}(\textit{{q}}_{1},\textit{{q}}_{2})(2\pi)^{3}\delta^{(D)}(\textit{{k}}-\textit{{q}}-\textit{{q}}_{1}-\textit{{q}}_{2}), (168)
A2\displaystyle A_{2} =\displaystyle= d3q(2π)3d3q1(2π)3d3q2(2π)3δ(1)(q)δ(1)(q1)δ(1)(q2)α~(q,kq)G2(q1,q2)(2π)3δ(D)(qq1q2).\displaystyle\int\frac{\text{d}^{3}q}{(2\pi)^{3}}\frac{\text{d}^{3}q_{1}}{(2\pi)^{3}}\frac{\text{d}^{3}q_{2}}{(2\pi)^{3}}\delta^{(1)}(\textit{{q}})\delta^{(1)}(\textit{{q}}_{1})\delta^{(1)}(\textit{{q}}_{2})\tilde{\alpha}(\textit{{q}},\textit{{k}}-\textit{{q}})\textit{G}_{2}(\textit{{q}}_{1},\textit{{q}}_{2})(2\pi)^{3}\delta^{(D)}(\textit{{q}}-\textit{{q}}_{1}-\textit{{q}}_{2}). (169)

and Eq. (166) as

Sβ~(3)(k,a)=2f2D3(a)(B1+B2),\begin{split}S^{(3)}_{\tilde{\beta}}(\textit{{k}},a)=-\mathcal{H}^{2}f^{2}D^{3}(a)(B_{1}+B_{2}),\end{split} (170)

where

B1\displaystyle B_{1} =\displaystyle= d3q(2π)3d3q1(2π)3d3q2(2π)3δ(1)(q)δ(1)(q1)δ(1)(q2)β~(q,kq)G2(q1,q2)(2π)3δ(D)(kqq1q2),\displaystyle\int\frac{\text{d}^{3}q}{(2\pi)^{3}}\frac{\text{d}^{3}q_{1}}{(2\pi)^{3}}\frac{\text{d}^{3}q_{2}}{(2\pi)^{3}}\delta^{(1)}(\textit{{q}})\delta^{(1)}(\textit{{q}}_{1})\delta^{(1)}(\textit{{q}}_{2})\tilde{\beta}(\textit{{q}},\textit{{k}}-\textit{{q}})\textit{G}_{2}(\textit{{q}}_{1},\textit{{q}}_{2})(2\pi)^{3}\delta^{(D)}(\textit{{k}}-\textit{{q}}-\textit{{q}}_{1}-\textit{{q}}_{2}), (171)
B2\displaystyle B_{2} =\displaystyle= d3q(2π)3d3q1(2π)3d3q2(2π)3δ(1)(q)δ(1)(q1)δ(1)(q2)β~(q,kq)G2(q1,q2)(2π)3δ(D)(qq1q2).\displaystyle\int\frac{\text{d}^{3}q}{(2\pi)^{3}}\frac{\text{d}^{3}q_{1}}{(2\pi)^{3}}\frac{\text{d}^{3}q_{2}}{(2\pi)^{3}}\delta^{(1)}(\textit{{q}})\delta^{(1)}(\textit{{q}}_{1})\delta^{(1)}(\textit{{q}}_{2})\tilde{\beta}(\textit{{q}},\textit{{k}}-\textit{{q}})\textit{G}_{2}(\textit{{q}}_{1},\textit{{q}}_{2})(2\pi)^{3}\delta^{(D)}(\textit{{q}}-\textit{{q}}_{1}-\textit{{q}}_{2}). (172)

The right-hand side of Eq. (164) can then be written as

Sβ~(3)a(aSα~(3))=2α2a3α(B1+B2)2a3α[12α+3α2](A1+A2),\displaystyle S^{(3)}_{\tilde{\beta}}-\mathcal{H}\partial_{a}\big{(}aS^{(3)}_{\tilde{\alpha}})=-\mathcal{H}^{2}\alpha^{2}a^{3\alpha}(B_{1}+B_{2})-\mathcal{H}^{2}a^{3\alpha}\Big{[}\frac{1}{2}\alpha+3\alpha^{2}\Big{]}(A_{1}+A_{2}),

from which we obtain

δ(3)(k)=12(8α+1)[2α(B1+B2)+(1+6α)(A1+A2)],\displaystyle\delta^{(3)}(\textbf{k})=\frac{1}{2(8\alpha+1)}\bigg{[}2\alpha(B_{1}+B_{2})+(1+6\alpha)(A_{1}+A_{2})\bigg{]},

which corresponds to

F3(q1,q2,q3)\displaystyle\textit{F}_{3}(\textit{{q}}_{1},\textit{{q}}_{2},\textit{{q}}_{3}) =\displaystyle= 12(8α+1)[(6α+1)α~(q1,q2+q3)×F2(q2,q3)+(2α)β~(q1,q2+q3)G2(q2,q3)\displaystyle\frac{1}{2(8\alpha+1)}\bigg{[}(6\alpha+1)\tilde{\alpha}(\textit{{q}}_{1},\textit{{q}}_{2}+\textit{{q}}_{3})\times\textit{F}_{2}(\textit{{q}}_{2},\textit{{q}}_{3})+(2\alpha)\tilde{\beta}(\textit{{q}}_{1},\textit{{q}}_{2}+\textit{{q}}_{3})\textit{G}_{2}(\textit{{q}}_{2},\textit{{q}}_{3})
+[(6α+1)α~(q1+q2,q3)+(2α)β~(q1+q2,q3)]G2(q1,q2)].\displaystyle+\big{[}(6\alpha+1)\tilde{\alpha}(\textit{{q}}_{1}+\textit{{q}}_{2},\textit{{q}}_{3})+(2\alpha)\tilde{\beta}(\textit{{q}}_{1}+\textit{{q}}_{2},\textit{{q}}_{3})\big{]}\textit{G}_{2}(\textit{{q}}_{1},\textit{{q}}_{2})\bigg{]}.

A.2 LPT

The left-hand side of Eq. (60) at third order can be written

(18α2+3α)2Ψi,i(3)(4α2+α)Ψi,j(1)Ψi,j(2)(2α2+α)2Ψi,j(2)Ψi,j(1)+(4α2+α)Ψi,i(1)Ψj,j(2)\displaystyle\frac{(18\alpha^{2}+3\alpha)}{2}\Psi_{i,i}^{(3)}-(4\alpha^{2}+\alpha)\Psi_{i,j}^{(1)}\Psi_{i,j}^{(2)}-\frac{(2\alpha^{2}+\alpha)}{2}\Psi_{i,j}^{(2)}\Psi_{i,j}^{(1)}+(4\alpha^{2}+\alpha)\Psi_{i,i}^{(1)}\Psi_{j,j}^{(2)}
(2α2+α)2Ψi,i(1)Ψi,j(1)Ψi,j(1)+(2α2+α)2Ψi,i(2)Ψi,j(1)+(2α2+α)4{Ψi,i(1)Ψj,j(1)Ψk,k(1)Ψi,i(1)Ψj,k(1)Ψk,j(1)}\displaystyle-\frac{(2\alpha^{2}+\alpha)}{2}\Psi_{i,i}^{(1)}\Psi_{i,j}^{(1)}\Psi_{i,j}^{(1)}+\frac{(2\alpha^{2}+\alpha)}{2}\Psi_{i,i}^{(2)}\Psi_{i,j}^{(1)}+\frac{(2\alpha^{2}+\alpha)}{4}\Big{\{}\Psi_{i,i}^{(1)}\Psi_{j,j}^{(1)}\Psi_{k,k}^{(1)}-\Psi_{i,i}^{(1)}\Psi_{j,k}^{(1)}\Psi_{k,j}^{(1)}\Big{\}}
+(2α2+α)2Ψi,l(1)Ψl,j(1)Ψi,j(1),\displaystyle+\frac{(2\alpha^{2}+\alpha)}{2}\Psi_{i,l}^{(1)}\Psi_{l,j}^{(1)}\Psi_{i,j}^{(1)}, (176)

and the right-hand side as

(2α2+α)2Ψi,i(3)+(2α2+α)2{Ψi,i(2)Ψj,j(1)Ψi,j(2)Ψj,i(1)}+(2α2+α)2{16Ψi,i(1)Ψj,j(1)Ψk,k(1)\displaystyle\frac{(2\alpha^{2}+\alpha)}{2}\Psi_{i,i}^{(3)}+\frac{(2\alpha^{2}+\alpha)}{2}\big{\{}\Psi_{i,i}^{(2)}\Psi_{j,j}^{(1)}-\Psi_{i,j}^{(2)}\Psi_{j,i}^{(1)}\big{\}}+\frac{(2\alpha^{2}+\alpha)}{2}\Big{\{}\frac{1}{6}\Psi_{i,i}^{(1)}\Psi_{j,j}^{(1)}\Psi_{k,k}^{(1)}
12Ψi,i(1)Ψj,k(1)Ψk,j(1)+13Ψi,l(1)Ψl,j(1)Ψi,j(1)}.\displaystyle-\frac{1}{2}\Psi_{i,i}^{(1)}\Psi_{j,k}^{(1)}\Psi_{k,j}^{(1)}+\frac{1}{3}\Psi_{i,l}^{(1)}\Psi_{l,j}^{(1)}\Psi_{i,j}^{(1)}\Big{\}}. (177)

Collecting up the terms we have

(8α+1)Ψi,i(3)\displaystyle(8\alpha+1)\Psi_{i,i}^{(3)} =\displaystyle= (4α+1)Ψi,i(2)Ψj,j(1)+(4α+1)Ψi,j(2)Ψj,i(1)+(2α+1)2Ψi,i(1)Ψi,j(1)Ψj,i(1)(2α+1)6Ψi,i(1)Ψj,j(1)Ψk,k(1)\displaystyle-(4\alpha+1)\Psi_{i,i}^{(2)}\Psi_{j,j}^{(1)}+(4\alpha+1)\Psi_{i,j}^{(2)}\Psi_{j,i}^{(1)}+\frac{(2\alpha+1)}{2}\Psi_{i,i}^{(1)}\Psi_{i,j}^{(1)}\Psi_{j,i}^{(1)}-\frac{(2\alpha+1)}{6}\Psi_{i,i}^{(1)}\Psi_{j,j}^{(1)}\Psi_{k,k}^{(1)} (178)
(2α+1)3Ψi,k(1)Ψk,j(1)Ψj,i(1),\displaystyle-\frac{(2\alpha+1)}{3}\Psi_{i,k}^{(1)}\Psi_{k,j}^{(1)}\Psi_{j,i}^{(1)},

which gives Eq. (62).

Appendix B Ultra-violet divergences in one-loop PS for gEdS

We detail here a little more the analysis leading to the expressions Eq. (83) and Eq. (84) which allows us to infer the ultraviolet convergence properties of these integrals i.e. the asymptotic large kk behaviour of Plin(k)P_{\rm lin}(k) for which they converge.

For the P22P_{22} the leading divergence is obtained simply by replacing Plin(|kq|)P_{\rm lin}(|\textbf{k}-\textbf{q}|) by Plin(q)P_{\rm lin}(q) and taking the leading term in the series expansion in k/qk/q around (k/q)=0(k/q)=0:

|F2(s)(kq,q)|2\displaystyle|F_{2}^{(s)}(\textbf{k}-\textbf{q},\textbf{q})|^{2} =\displaystyle= 14k4q4+(4α+16α+1)[(μ21)k4q4]+(4α+16α+1)2[(μ21)2k4q4]+\displaystyle\frac{1}{4}\frac{k^{4}}{q^{4}}+\Big{(}\frac{4\alpha+1}{6\alpha+1}\Big{)}\Big{[}(\mu^{2}-1)\frac{k^{4}}{q^{4}}\Big{]}+\Big{(}\frac{4\alpha+1}{6\alpha+1}\Big{)}^{2}\Big{[}(\mu^{2}-1)^{2}\frac{k^{4}}{q^{4}}\Big{]}+\dots
=\displaystyle= (2α(8α+2)μ2+12(6α+1))2k4q4+\displaystyle\Big{(}\frac{2\alpha-(8\alpha+2)\mu^{2}+1}{2(6\alpha+1)}\Big{)}^{2}\frac{k^{4}}{q^{4}}+\cdots

where μ=kqkq\mu=\frac{\textbf{k}\cdot\textbf{q}}{kq} is the angular variable. Using Eqs. (73) and (B) we then obtain, after integrating μ\mu, Eq. (83).

Likewise expanding 11𝑑μF3(s)(k,q,q)\int_{-1}^{1}d\mu F_{3}^{(s)}(\textbf{k},\textbf{q},-\textbf{q}) in a power series in k/qk/q around (k/q)=0(k/q)=0 we obtain

limk/q011𝑑μF3(s)(k,q,q)\displaystyle\lim_{k/q\to 0}\int_{-1}^{1}d\mu F_{3}^{(s)}(\textbf{k},\textbf{q},-\textbf{q}) =\displaystyle= 19k2q2+(2α+18α+1)[4105k4q4+×415k2q2]+(2α+18α+1)(2α6α+1)[415k4q449k2q2]+\displaystyle-\frac{1}{9}\frac{k^{2}}{q^{2}}+\Big{(}\frac{2\alpha+1}{8\alpha+1}\Big{)}\Big{[}-\frac{4}{105}\frac{k^{4}}{q^{4}}+\times\frac{4}{15}\frac{k^{2}}{q^{2}}\Big{]}+\Big{(}\frac{2\alpha+1}{8\alpha+1}\Big{)}\Big{(}\frac{2\alpha}{6\alpha+1}\Big{)}\Big{[}\frac{4}{15}\frac{k^{4}}{q^{4}}-\frac{4}{9}\frac{k^{2}}{q^{2}}\Big{]}+\dots (180)
=\displaystyle= 714α176α245(1+6α)(1+8α)k2q2+4(8α1)(1+2α)105(1+6α)(1+8α)k4q4+\displaystyle\frac{7-14\alpha-176\alpha^{2}}{45(1+6\alpha)(1+8\alpha)}\frac{k^{2}}{q^{2}}+\frac{4(8\alpha-1)(1+2\alpha)}{105(1+6\alpha)(1+8\alpha)}\frac{k^{4}}{q^{4}}+\dots

which leads directly to Eq. (84).

Appendix C Summary of calculation of cosmology dependent corrections to the EdS approximation for the one loop PS in standard perturbation theory

We have shown that this correction is given by

ΔP1loop=249γ2(M1+107M2)+(249γ2)2M2+449[γ2N1+554γ3(N2N1)],\displaystyle\Delta P_{1-loop}=\frac{2}{49}\gamma_{2}(M_{1}+\frac{10}{7}M_{2})+(\frac{2}{49}\gamma_{2})^{2}M_{2}+\frac{4}{49}\big{[}\gamma_{2}N_{1}+\frac{5}{54}\gamma_{3}(N_{2}-N_{1})\big{]}, (181)

where the four integrals are given in Eqs. (77) and (81). These expressions are valid for a generic FLRW cosmology with a clustering matter component and a smooth component that scales at asymptotically early times as matter i.e. with w0w\rightarrow 0 as a0a\rightarrow 0 (where w=13dΩm(a)dlnaw=\frac{1}{3}\frac{d\Omega_{m}(a)}{d\ln a}). The two functions γ2(a)\gamma_{2}(a) and γ3(a)\gamma_{3}(a) are obtained by solving the coupled equations:

dα1dlna\displaystyle\frac{d\alpha_{1}}{d\ln a} =\displaystyle= 32wα1α1212α1+32Ωm,\displaystyle\frac{3}{2}w\alpha_{1}-\alpha_{1}^{2}-\frac{1}{2}\alpha_{1}+\frac{3}{2}\Omega_{m}, (182)
d2γ2dlna2\displaystyle\frac{d^{2}\gamma_{2}}{d\ln a^{2}} =\displaystyle= [12(13w)+4α1]dγ2dlna(2α12+32Ωm)γ2+212(Ωmα12),\displaystyle-\Big{[}\frac{1}{2}(1-3w)+4\alpha_{1}\Big{]}\frac{d\gamma_{2}}{d\ln a}-\Big{(}2\alpha_{1}^{2}+\frac{3}{2}\Omega_{m}\Big{)}\gamma_{2}+\frac{21}{2}(\Omega_{m}-\alpha_{1}^{2}), (183)
d2γ3dlna2\displaystyle\frac{d^{2}\gamma_{3}}{d\ln a^{2}} =\displaystyle= [12(13w)+6α1]dγ3dlna3(2α12+Ωm)γ31895(Ωmα12)+18935[4α12γ2+2α1dγ2dlna],\displaystyle-\Big{[}\frac{1}{2}(1-3w)+6\alpha_{1}\Big{]}\frac{d\gamma_{3}}{d\ln a}-3(2\alpha_{1}^{2}+\Omega_{m})\gamma_{3}-\frac{189}{5}\Big{(}\Omega_{m}-\alpha_{1}^{2}\Big{)}+\frac{189}{35}\Big{[}4\alpha_{1}^{2}\gamma_{2}+2\alpha_{1}\frac{d\gamma_{2}}{d\ln a}\Big{]}, (184)

subject to the boundary conditions, given at a=0a=0, by

α1(0)=α2(0)=α3(0)=α10(0),dα2da(0)=dα3da(0)=0,\alpha_{1}(0)=\alpha_{2}(0)=\alpha_{3}(0)=\alpha_{10}(0),\,\frac{d\alpha_{2}}{da}(0)=\frac{d\alpha_{3}}{da}(0)=0, (185)

or, more explicitly,

γ2(0)\displaystyle\gamma_{2}(0) =\displaystyle= 71α10(0)1+6α10(0),\displaystyle 7\frac{1-\alpha_{10}(0)}{1+6\alpha_{10}(0)},
γ3(0)\displaystyle\gamma_{3}(0) =\displaystyle= 635(1α10(0))(6α10(0)1)(1+6α10(0))1+8α10(0)),\displaystyle\frac{63}{5}\frac{(1-\alpha_{10}(0))(6\alpha_{10}(0)-1)}{(1+6\alpha_{10}(0))1+8\alpha_{10}(0))}, (186)

and

dγ2da(0)=0,dγ3da(0)=0,\frac{d\gamma_{2}}{da}(0)=0\,\,,\frac{d\gamma_{3}}{da}(0)=0, (187)

and where α10(0)\alpha_{10}(0) is given by Eq. (91) with Ωm\Omega_{m} given by the asymptotic (constant) fraction of clustering matter at high redshift. Given that α1=α2=α3(0)=α10(a)\alpha_{1}=\alpha_{2}=\alpha_{3}(0)=\alpha_{10}(a) is the solution at leading order in ww, we can integrate in practice from 0<ai10<a_{i}\ll 1 assuming exactly the initial conditions given by this solution. Note that in a standard EdS cosmology we have of course α10(0)=1\alpha_{10}(0)=1, and the case α10(0)1\alpha_{10}(0)\neq 1 corresponds to the more general gEdS case where part of the matter component at high redshift is smooth (e.g. a massive neutrino component or matter-like dark energy component).

Appendix D Detailed comparison with results of [27]

Expressions for the cosmological corrections, at second and third order in EPT, and for the one-loop PS (as in the previous appendix), have been given in [27], in terms of six time-dependent functions. Further this paper provides phenomenological fits for the parameter dependence in the case of a smooth component with a constant equation of state. We make explicit here how these six functions are related to one another by the three additional constraints we have derived (which reduces the number of independent functions to three, of which only two are required to calculate the PS at one loop). We also check that there is indeed good numerical agreement between our results and those of [27].

D.1 Second-order growth factor

[27] gives results for the second-order growth factor in terms of two numerically fitted functions as

T2A\displaystyle T_{2A} =\displaystyle= D2AD121=|lnΩm|(5.54×103|w0|3.40×103|w0|),\displaystyle\frac{D_{2A}}{D_{1}^{2}}-1=|\ln{\Omega_{m}}|\Big{(}\frac{5.54\times 10^{-3}}{|w_{0}|}-\frac{3.40\times 10^{-3}}{\sqrt{|w_{0}|}}\Big{)}, (188)
T2B\displaystyle T_{2B} =\displaystyle= D2BD121=|lnΩm|(1.384×102|w0|+8.50×103|w0|).\displaystyle\frac{D_{2B}}{D_{1}^{2}}-1=|\ln{\Omega_{m}}|\Big{(}-\frac{1.384\times 10^{-2}}{|w_{0}|}+\frac{8.50\times 10^{-3}}{\sqrt{|w_{0}|}}\Big{)}. (189)

These are related (cf. Eqs. (108)) to our effective exponent γ2\gamma_{2} by

T2A=235γ2,T2B=17γ2,T_{2A}=\frac{2}{35}\gamma_{2},\qquad T_{2B}=-\frac{1}{7}\gamma_{2}, (190)

and thus T2A=(2/5)T2BT_{2A}=-(2/5)T_{2B}. Comparing the coefficients we see that this relation is indeed verified within the level of numerical precision quoted by [27].444What we have denoted by w0w_{0} here corresponds to ww in [27].

Refer to caption
Figure 9: The ratio of our numerical solutions for γ2\gamma_{2} to the parametric fit of [27] γ2,Tak\gamma_{2,Tak} for the same quantity, for dark energy with constant equation of state w0=0.5,1.0w_{0}=-0.5,-1.0, and 1.5-1.5.
Refer to caption
Refer to caption
Figure 10: Left panel: Ratio of the function γ3AA\gamma_{3AA} given by our direct numerical solution of Eq. (142) to its value inferred from the fits of [27], for a smooth component with constant equation of state (w0=0.5,1w_{0}=-0.5,-1 and 1.5-1.5). Right panel: Same comparison, but only for the model with w0=1w_{0}=-1, and for γ3=γ3BA\gamma_{3}=\gamma_{3BA} (from numerical solution of Eq. (184)). The two different curves correspond to the two different possible numerical fits which can be inferred from those of [27]: the full line corresponds to that inferred from T3BAT_{3BA} directly, and the dashed line to that obtained using the linear combination on the right-hand side of Eq. (200)). The level of accuracy is less good than for γ2\gamma_{2} and γ3AA\gamma_{3AA}, but quite adequate for calculation notably of the PS at any practically relevant level of precision.

Fig. 9 shows the ratio between the results we obtain for γ2\gamma_{2} by numerical integration of Eqs.(182) and (183) with the fit for the same quantity inferred from [27]:

γ2,Tak=(35/2)T2A|lnΩm|(9.7×102|w0|5.95×102|w0|),\displaystyle\gamma_{2,Tak}=(35/2)T_{2A}\approx|\ln{\Omega_{m}}|\Big{(}\frac{9.7\times 10^{-2}}{|w_{0}|}-\frac{5.95\times 10^{-2}}{\sqrt{|w_{0}|}}\Big{)},

We see that our numerical solution for γ2\gamma_{2} well agree very accurately with Eq. (D.1) for w0=0.5,1w_{0}=-0.5,-1 and well, albeit slightly less so, for w0=1.5w_{0}=-1.5.

D.2 Third-order growth factor

At this order [27] gives results in terms of the four functions

T3AA=D3AAD131\displaystyle T_{3AA}=\frac{D_{3AA}}{D_{1}^{3}}-1 \displaystyle\simeq |lnΩm|(8.21×103|w0|5.14×103|w0|),\displaystyle|\ln{\Omega_{m}}|\Big{(}\frac{8.21\times 10^{-3}}{|w_{0}|}-\frac{5.14\times 10^{-3}}{\sqrt{|w_{0}|}}\Big{)}, (192)
T3AB=D3ABD131\displaystyle T_{3AB}=\frac{D_{3AB}}{D_{1}^{3}}-1 \displaystyle\simeq |lnΩm|1.5+0.4ln|w0||Ωm|0.7|w0|(9.16×103|w0|+8.95×103|w0|),\displaystyle|\ln{\Omega_{m}}|^{1.5+0.4\ln{|w_{0}|}}|\Omega_{m}|^{0.7|w_{0}|}\Big{(}-\frac{9.16\times 10^{-3}}{|w_{0}|}+\frac{8.95\times 10^{-3}}{\sqrt{|w_{0}|}}\Big{)}, (193)
T3BA=D3BAD131\displaystyle T_{3BA}=\frac{D_{3BA}}{D_{1}^{3}}-1 \displaystyle\simeq |lnΩm|1.060.5ln|w0|(7.68×103|w0|1.130×102|w0|),\displaystyle|\ln{\Omega_{m}}|^{1.06-0.5\ln{|w_{0}|}}\Big{(}7.68\times 10^{-3}|w_{0}|-1.130\times 10^{-2}\sqrt{|w_{0}|}\Big{)}, (194)
T3BB=D3BBD131\displaystyle T_{3BB}=\frac{D_{3BB}}{D_{1}^{3}}-1 \displaystyle\simeq |lnΩm|(2.641×102|w0|+1.582×102|w0|),\displaystyle|\ln{\Omega_{m}}|\Big{(}-\frac{2.641\times 10^{-2}}{|w_{0}|}+\frac{1.582\times 10^{-2}}{\sqrt{|w_{0}|}}\Big{)}, (195)

and two additional functions D3AAD_{3AA}^{\prime} and D3ABD_{3AB}^{\prime} determined by the relations in Eqs. (IV.3).

These are related to the functions we have introduced by

T3AA\displaystyle T_{3AA} =\displaystyle= 445γ3AA,T3AB=19γ3AB,\displaystyle\frac{4}{45}\gamma_{3AA},\quad T_{3AB}=-\frac{1}{9}\gamma_{3AB},
T3BA\displaystyle T_{3BA} =\displaystyle= 563γ3BA,T3BB=1663γ3BB.\displaystyle\frac{5}{63}\gamma_{3BA},\quad T_{3BB}=-\frac{16}{63}\gamma_{3BB}. (196)

From Eqs. (130)-(131) it is straightforward to show that our additional constraints relating to these functions are

98γ3AA49γ3AB+15γ3BA64γ3BB=0,\displaystyle 98\gamma_{3AA}-49\gamma_{3AB}+15\gamma_{3BA}-64\gamma_{3BB}=0, (197)
49γ3AA+32γ3BB=81γ2.\displaystyle 49\gamma_{3AA}+32\gamma_{3BB}=81\gamma_{2}. (198)

Using these we can infer the relations

T3BB\displaystyle T_{3BB} =\displaystyle= 92T2B+358T3AA,\displaystyle\frac{9}{2}T_{2B}+\frac{35}{8}T_{3AA}, (199)
T3BA\displaystyle T_{3BA} =\displaystyle= (43T3BB+356T3AA+73T3AB).\displaystyle-\big{(}\frac{4}{3}T_{3BB}+\frac{35}{6}T_{3AA}+\frac{7}{3}T_{3AB}\big{)}. (200)

For the first relation it is easy to check the accuracy with which it is satisfied by the expressions above from [27], because the same functional form has been used to fit the functions on both sides: comparing the two fitted coefficients on the left and right-hand side of the relation Eq. (199), we find they agree at the sub-percent level. Further to check the numerical accuracy of these fits against ours we plot in the left panel of Fig. 10 the ratio between our own numerical solution for γ3AA\gamma_{3AA} and γ3AA,Tak=(45/4)T3AA\gamma_{3AA,Tak}=(45/4)T_{3AA}. The agreement is at a comparable level to that for γ2\gamma_{2}.

To assess both the agreement of Eq. (200) with the fits of [27] given above, and the accuracy of these fits in describing the exact solutions, we plot in the right panel of Fig. 10 the ratio of our numerical solutions for γ3=γ3BA\gamma_{3}=\gamma_{3BA} to γ3BA,Tak1=(63/5)T3BA\gamma_{3BA,Tak1}=(63/5)T_{3BA} and γ3BA,Tak2=(63/5)T3BA\gamma_{3BA,Tak2}=(63/5)T_{3BA}^{\prime} respectively, where T3BAT_{3BA}^{\prime} is the linear combination on the right-hand side of Eq. (200). These curves are for a model with w0=1w_{0}=-1. We can infer that Eq. (200) is obeyed by the fitting formulae at the 10%10\% level, and that the agreement with our numerical results is slightly better if we use the fit for T3BAT_{3BA}. We have checked carefully the origin of these small discrepancies and find that they arise from transients from initial conditions which disappear. In practice however, notably for the PS corrections which depend only very weakly on γ3\gamma_{3}, the fits provided by [27] are accurate enough for any current practical application.

Appendix E Comparison with analysis of [32]

Refer to caption
Refer to caption
Figure 11: Our numerical solutions for γ2\gamma_{2}, and γ3=γ3BA\gamma_{3}=\gamma_{3BA} (solid black lines) in LCDM (w0=1w_{0}=-1) compared with the power series expressions in Eqs. (203)-(204) obtained from [32]. The different lines correspond to including terms in the power series up to the indicated order.

In [32] the time dependence of the second order kernel is parametrized by two functions λ2(1)\lambda_{2}^{(1)} and λ2(2)\lambda_{2}^{(2)}, which are identical to the two functions we have used:

λ2(1)=d2A,λ2(2)=d2B.\displaystyle\lambda_{2}^{(1)}=d_{2A},\quad\lambda_{2}^{(2)}=d_{2B}. (201)

The six functions used to parametrize the third-order density kernels just differ by a factor of 22:

λ3(1)=12d3AA,λ3(2)=12d3AB,\displaystyle\lambda_{3}^{(1)}=\frac{1}{2}d_{3AA},\quad\lambda_{3}^{(2)}=\frac{1}{2}d_{3AB},
λ3(3)=12d3AA,λ3(4)=12d3AB,\displaystyle\lambda_{3}^{(3)}=\frac{1}{2}d_{3AA}^{\prime},\quad\lambda_{3}^{(4)}=\frac{1}{2}d_{3AB}^{\prime},
λ3(5)=12d3BA,λ3(6)=12d3BB.\displaystyle\lambda_{3}^{(5)}=\frac{1}{2}d_{3BA},\quad\lambda_{3}^{(6)}=\frac{1}{2}d_{3BB}. (202)

For the specific case of a LCDMLCDM cosmology, [32] provide analytical solutions for these eight functions (and further ones parametrizing the fourth and fifth order kernels) as power series in the parameter ζ=ΩΛ0Ωm0e3η\zeta=\frac{\Omega_{\Lambda 0}}{\Omega_{m0}}e^{3\eta}, where ηlnD+\eta\equiv\ln D_{+} (D+D_{+} the linear growth factor normalized to unity at z=0z=0, and ΩΛ0\Omega_{\Lambda 0} and Ωm0\Omega_{m0} are the dark energy and matter fractions at z=0z=0, respectively). It is straightforward to verify that the expressions given in Appendix A of [32], up to third order in η\eta, satisfy the five relations Eq. (106) and Eqs. (137)-(140).

To check the numerical agreement of our results with those of [32] we calculate the functions we have used to parametrize the PS and find

γ2\displaystyle\gamma_{2} =\displaystyle= 72(7λ2(1)5)=7c126ζ2c219ζ27c3125ζ3,\displaystyle\frac{7}{2}(7\lambda_{2}^{(1)}-5)=-\frac{7c_{1}}{26}\zeta-\frac{2c_{2}}{19}\zeta^{2}-\frac{7c_{3}}{125}\zeta^{3}, (203)
γ3BA\displaystyle\gamma_{3BA} =\displaystyle= 635(21λ3(5)1)=462c11625ζ+51c2266ζ2+2576c320625ζ3,\displaystyle\frac{63}{5}(21\lambda_{3}^{(5)}-1)=\frac{462c_{1}}{1625}\zeta+\frac{51c_{2}}{266}\zeta^{2}+\frac{2576c_{3}}{20625}\zeta^{3}, (204)

where

c1=332,c2=1414114,c3=99931040842c_{1}=-\frac{3}{32},\quad c_{2}=-\frac{141}{4114},\quad c_{3}=-\frac{9993}{1040842} (205)

Figure 11 shows the comparison between our numerical solutions to the exact equations for the functions γ2\gamma_{2} and γ3=γ3BA\gamma_{3}=\gamma_{3BA} and the power series solutions of [32] for the same quantities, in LCDM models, up to the indicated orders. We observe excellent agreement where the power series solutions appear to converge well.

Appendix F Comparison with analysis of [11, 20]

[11] parametrized the time dependence of second order and third order density kernels in terms of three functions. Comparing their definitions directly to ours (equivalent to those of [27, 32]) we obtain

d2A\displaystyle d_{2A} =\displaystyle= 12+34ν2,\displaystyle-\frac{1}{2}+\frac{3}{4}\nu_{2}, (206)
d2B\displaystyle d_{2B} =\displaystyle= 3234ν2,\displaystyle\frac{3}{2}-\frac{3}{4}\nu_{2}, (207)
d3AA\displaystyle d_{3AA} =\displaystyle= λ323ν24+3ν38+12,\displaystyle-\frac{\lambda_{3}}{2}-\frac{3\nu_{2}}{4}+\frac{3\nu_{3}}{8}+\frac{1}{2}, (208)
d3AA\displaystyle d_{3AA}^{\prime} =\displaystyle= λ363ν24+3ν38+16,\displaystyle\frac{\lambda_{3}}{6}-\frac{3\nu_{2}}{4}+\frac{3\nu_{3}}{8}+\frac{1}{6}, (209)
d3AB\displaystyle d_{3AB} =\displaystyle= 7λ36+3ν243ν38+16,\displaystyle\frac{7\lambda_{3}}{6}+\frac{3\nu_{2}}{4}-\frac{3\nu_{3}}{8}+\frac{1}{6}, (210)
d3AB\displaystyle d_{3AB}^{\prime} =\displaystyle= λ32+3ν243ν38+12,\displaystyle\frac{\lambda_{3}}{2}+\frac{3\nu_{2}}{4}-\frac{3\nu_{3}}{8}+\frac{1}{2}, (211)
d3BA\displaystyle d_{3BA} =\displaystyle= λ36+9ν243ν38136,\displaystyle-\frac{\lambda_{3}}{6}+\frac{9\nu_{2}}{4}-\frac{3\nu_{3}}{8}-\frac{13}{6}, (212)
d3BB\displaystyle d_{3BB} =\displaystyle= λ329ν24+3ν38+52.\displaystyle-\frac{\lambda_{3}}{2}-\frac{9\nu_{2}}{4}+\frac{3\nu_{3}}{8}+\frac{5}{2}. (213)

It is straightforward to verify that, using these expressions, the five relations Eq. (106) and Eqs. (137)-(140) are satisfied. As we have shown in Section IV, this corresponds to assuming that the eight functions obey conditions valid for EdS boundary conditions, as assumed by [11]. As we have seen in Section IV, these same conditions are also appropriate for the more general case of gEdS boundary conditions.

The three effective growth rate functions (α2\alpha_{2}, α3AA\alpha_{3AA}, and α3AB\alpha_{3AB}) that we have chosen to characterize the second order and third order kernels can be inferred from

ν2\displaystyle\nu_{2} =\displaystyle= 23(2d2A+1),\displaystyle\frac{2}{3}\left(2d_{2A}+1\right), (214)
ν3\displaystyle\nu_{3} =\displaystyle= 23(4d2A+7d3AA+3d3AB2),\displaystyle\frac{2}{3}(4d_{2A}+7d_{3AA}+3d_{3AB}-2), (215)
λ3\displaystyle\lambda_{3} =\displaystyle= 12(3d3AA+3d3AB2).\displaystyle\frac{1}{2}(3d_{3AA}+3d_{3AB}-2). (216)

by substituting the appropriate definitions, Eq. (107) and Eqs. (LABEL:def-gammas-3). In terms of the functions γ2\gamma_{2}, γ3AA\gamma_{3AA}, and γ3AB\gamma_{3AB} we have

ν2\displaystyle\nu_{2} =\displaystyle= 2147(119+4γ2),\displaystyle\frac{2}{147}(119+4\gamma_{2}), (217)
ν3\displaystyle\nu_{3} =\displaystyle= 211907(1372γ3AA294γ3AB+648γ2+21483),\displaystyle\frac{2}{11907}(1372\gamma_{3AA}-294\gamma_{3AB}+648\gamma_{2}+21483), (218)
λ3\displaystyle\lambda_{3} =\displaystyle= 154(4γ3AA2γ3AB+9).\displaystyle\frac{1}{54}(4\gamma_{3AA}-2\gamma_{3AB}+9). (219)

[20] give (Appendix A) numerical values for ν2\nu_{2}, ν3\nu_{3} and λ3\lambda_{3} in an LCDM model with Ωm=0.3\Omega_{m}=0.3 at z=0z=0. For this case our numerical integration gives γ2=0.0452\gamma_{2}=0.0452\dots, γ3AA=0.0417\gamma_{3AA}=0.0417\dots and γ3AB=0.00142\gamma_{3AB}=0.00142\dots. Using Eqs. (217)-(219) we find numerical agreement up to at least four significant decimal places.

We note also that our result, that the one loop PS can be written in terms of d2Ad_{2A} and d3BAd_{3BA} only, implies that it can equivalently be written only in terms of ν2\nu_{2} and the linear combination (4λ3+9ν3)(4\lambda_{3}+9\nu_{3}). Finally we note that in the gEdS limit, corresponding to α2=α3AA=α3AB=α\alpha_{2}=\alpha_{3AA}=\alpha_{3AB}=\alpha we obtain

ν2\displaystyle\nu_{2} =\displaystyle= 23(3+14α1+6α),\displaystyle\frac{2}{3}\Big{(}\frac{3+14\alpha}{1+6\alpha}\Big{)}, (220)
ν3\displaystyle\nu_{3} =\displaystyle= 23(236α2+96α+9)(1+6α)(1+8α),\displaystyle\frac{2}{3}\frac{\left(236\alpha^{2}+96\alpha+9\right)}{(1+6\alpha)(1+8\alpha)}, (221)
λ3\displaystyle\lambda_{3} =\displaystyle= 12(1+2α1+8α).\displaystyle\frac{1}{2}\Big{(}\frac{1+2\alpha}{1+8\alpha}\Big{)}. (222)

References