<<

Newtonian Versus Relativistic Cosmology

Master Thesis

by

Samuel Flender

1st corrector: Prof. Dr. Dominik Schwarz 2nd corrector: Prof. Dr. Dietrich Bödeker

University of Bielefeld January 2012 Contents

1 Introduction 4

2 The homogeneous and isotropic universe 6 2.1 Basic equations ...... 6 2.2 Content of the universe ...... 7 2.3 Ination ...... 8 2.4 Evolution of scales ...... 10

3 The perturbed universe: Newtonian treatment 12 3.1 Denitions ...... 12 3.2 Perturbed elds ...... 12 3.3 Perturbed Newtonian equations ...... 13 3.3.1 Continuity equation ...... 13 3.3.2 Poisson equation ...... 14 3.3.3 Euler equation ...... 15 3.3.4 Summary ...... 17 3.4 Solutions ...... 17 3.4.1 First order solutions ...... 17 3.4.2 Initial conditions ...... 19 3.4.3 Transfer function ...... 20 3.4.4 Second order solutions ...... 24 3.4.5 Discussion ...... 28 3.4.6 Solutions during Λ-domination ...... 31 3.5 Newtonian cosmological simulations ...... 31

4 The perturbed universe: relativistic treatment 32 4.1 The perturbed metric ...... 32 4.2 Gauge transformations ...... 33 4.2.1 Gauge transformations of scalars, vectors and ...... 35 4.2.2 Gauge transformation of metric perturbations ...... 36 4.2.3 Gauge transformations of velocity and density perturbations ...... 37 4.2.4 Gauge-invariant quantities ...... 38 4.2.5 Gauges ...... 39 4.3 Relativistic equations in the conformal Newtonian gauge ...... 42 4.4 Solutions ...... 44 4.4.1 Solutions in the conformal Newtonian gauge ...... 44 4.4.2 Solutions in other gauges ...... 45 4.4.3 Gauge-invariant solutions ...... 46 4.5 Vector and perturbations ...... 48

2 5 Quantication of relativistic corrections 49 5.1 Relativistic corrections to Newtonian simulations ...... 49 5.2 Relativistic corrections to the observed power spectrum ...... 51 5.3 Relativistic corrections in ΛCDM models ...... 52

6 Summary and outlook 57

Appendix 58 Appendix A: Helmholtz decomposition theorem ...... 58 Appendix B: Statistical properties of Gaussian perturbations ...... 60 Appendix C: Perturbations in the inaton eld ...... 62 Appendix D: Meszaros equation ...... 63 Appendix E: Identity of synchronous and comoving gauge ...... 64

References 66

Acknowledgments 68

3 Figure 1: The large scale structure of the universe. This map was created using data from the Sloan digital sky survey. Each dot represents a galaxy. [source: http://www.sdss.org/]

1 Introduction

Cosmology is the study of the origin and structure of the universe. We have good observational ev- idence that the universe was isotropic and homogeneous at early times, the best evidence being the small uctuations (about ∼ 10−5) of the cosmic microwave background (CMB) radiation [1]. However, the structure of the universe that we observe today is very inhomogeneous on scales below ∼ 100 Mpc1: there are galaxies (on scales of ∼ 1 Mpc) and clusters of galaxies (on scales of ∼ 10 Mpc). On scales of ∼ 100 Mpc there are superclusters of galaxies, which are separated by large voids [2], see gure 1. The theory of connects the early homogeneous universe with the inhomogeneous universe today: It is assumed that there exist initially small density perturbations, which grow due to gravitational attraction. The mathematical tool to describe this scenario accurately is relativistic cosmological perturbation theory, where one considers small perturbations on the homogeneous and isotropic Friedmann-Robertson-Walker (FRW) metric. However, in this theory there are complica- tions arising due to the freedom of gauge, i.e. choosing the correspondence between perturbed and background quantities. Another approach is Newtonian cosmological perturbation theory, which results in perturbed New- tonian equations (continuity equation, Euler equation and Poisson equation) on a FRW background.

1Distances in cosmology are usually measured in units of Millions of parsec (Megaparsec, Mpc), where one parsec (pc) is roughly 3.26 lightyears.

4 Figure 2: Result of the Millennium simulation. [source: http://www.mpa-garching.mpg.de]

One has to keep in mind that Newtonian theory is wrong in that it assumes instantaneous gravita- tional interaction and innite speed of light. However, it turns out that Newtonian and relativistic cosmological perturbation theory coincide for scales much smaller than the Hubble horizon, which is basically the size of the observable universe.

Cosmological N-body simulations use Newtonian cosmology to move particles and simulate the density growth in the universe. These simulations can be used to test our understanding of structure formation. A famous simulation is the Millennium run (see gure 2), which simulated the behaviour of ∼ 1010 particles (each particle representing a dark matter clump with a mass of 109 suns) in a box of ∼ 1Gpc side length [3]. But the question is, how reliable are these simulations, since they use Newtonian rather than relativistic equations. In this work, we want to study the relativistic corrections to the Newtonian equations. Cosmological perturbation theory is a major tool to do this.

The outline is as follows: In part 2, we will summarize our current understanding of the universe and give the most important equations describing it. In part 3, we are going to consider Newtonian perturbation theory on an expanding FRW background. We will derive and solve the perturbed New- tonian equations up to second order. In part 4, we will introduce relativistic cosmological perturbation theory and derive and solve the relativistic evolution equations for the perturbation variables. In part 5, we will quantify the relativistic corrections to cosmological simulations and the observed matter power spectrum.

5 2 The homogeneous and isotropic universe

2.1 Basic equations

This section resumes the basic equations in cosmology. For a reference, see e.g. [4, 5]. We set c = 1. The Friedmann equation relates the Hubble parameter to the content of the universe:

8πG k H2 = ρ(t) − , (2.1) 3 a2 where is the scale factor, 1 ∂a a˙ is the Hubble parameter, is the curvature of the universe a H ≡ a ∂t = a k and ρ(t) is its energy density. This can consist of matter density, radiation density, or something else. Dene the critical energy density 3H2 ρ ≡ (2.2) c 8πG and the relative energy density ρ(t) Ω(t) ≡ . (2.3) ρc(t) Then we can rewrite the Friedmann equation as

k 1 − Ω(t) = − . (2.4) a2H2

From this equation it can be seen that the universe has critical energy density (Ω = 1) if and only if the curvature k vanishes. Furthermore, an overcritical universe (Ω > 1) has positive curvature and an undercritical universe (Ω < 1) has negative curvature. The equation of state connects the energy density of a given source of energy and the pressure P that is created by it, P = wρ, (2.5) where w is the equation of state parameter. Another important equation is the continuity equation, which tells how much the energy density will change in time, ρ˙ + 3H(ρ + P ) = 0. (2.6)

Inserting the equation of state, we nd

ρ˙ + 3Hρ(1 + w) = 0, (2.7) which has the solution −3(1+w) ρ = ρ0a , (2.8) where ρ0 = ρ(t0) is the energy density today (we scale a such that a(t0) = 1). For a spatially at single uid universe, the Friedmann equation then becomes

8πG a˙ 2 = ρ a−(1+3w). (2.9) 3 0

6 Figure 3: Content of the universe today. [source: map.gsfc.nasa.gov/universe/uni_matter.html]

Given an equation of state parameter w we can solve this dierential equation for the scale factor a(t). Finally, we introduce the Raychaudhuri equation, which can be derived from the Friedmann equation and the continuity equation. It gives the acceleration of the scale factor in terms of energy density and pressure, a¨ 4πG = − (ρ + 3P ). (2.10) a 3

2.2 Content of the universe

Now we discuss the dierent components of our universe. The universe contains:

• Matter. This is to a small part (Ωb,0 ' 0.05) baryonic matter, which is mostly hydrogen. Everything that we see, stars, galaxies and clusters of galaxies, but also interstellar dust is contained in this number. However, various observations suggest that there is another type of matter that we cannot see directly, but its gravitational eects [6]. We call this type of matter −3 dark matter. Matter density ρm, dened as mass per volume, scales as ρm(t) ∼ a , because 3 volume scales as a . Thus, the equation of state parameter is wm = 0. In other words, matter does not create any pressure. Today, baryonic and dark matter together make up a relative

energy density of Ωm,0 ' 0.28, but the biggest part of it comes from dark matter, Ωdm,0 ' 0.23.

−4 • Radiation. Radiation density scales as ργ ∼ a , because the number density of photons scales as −3 and the energy of each photon scales as 1 −1. Hence, the equation of nγ ∼ a Eγ ∼ λ ∼ a state parameter is 1 . Measurements show that today −4, which comes mainly wγ = 3 Ωγ,0 ' 10 from CMB (cosmic microwave background) photons.

• Curvature. It is not known for sure if the universe is spatially at. If curvature exists, then it acts −2 eectively as an energy density, as can be seen from the Friedmann equation, with ρk ∼ a , so that 1 . However, measurements show that the eect of curvature cannot be large today, wk = − 3 −2. |Ωk,0| . 10

7 

 ~ −4  a

−  ~a 3 m  =  const

radiation domination matter domination dark energy domination today a

Figure 4: Overview over the dierent epochs of our universe (in a logarithmic scaling). Because radiation, matter and dark energy density scale dierently with time, there exist eras where one of them is dominating.

• Λ. Dark energy (also called vacuum energy or cosmological constant Λ) is a component of the universe whose energy density does not change with time. Consequently, the equation of state

parameter is wΛ = −1, whence it creates negative pressure. Measurements of Type IA supernovae show that the universe today is expanding with an accelerating scale factor. These measurements can be explained by the existence of a cosmological constant with a relative energy density of

ΩΛ,0 ' 0.72. Thus, the biggest contribution to the energy density of the universe today comes from dark energy [7].

Now that we know how the dierent components of the universe scale, we can give an approximation when which component was dominant. Current cosmological models say that after the big bang, there was a short time when the universe was dark energy dominated, so that it expanded exponentially. This period is called ination. It broke down when the slow-roll conditions were violated, to be explained below. After this, the universe got radiation-dominated (a ∼ t1/2) until the time of radiation-matter Ωγ,0 −4 2/3 equality at aeq = ' 3 · 10 . Then it became matter-dominated (a ∼ t ) until the time of Ωm,0 1/3  Ωm,0  matter-Λ equality at amΛ = ' 0.7. Thus, today (a0 = 1) we live at a time just after the ΩΛ,0 cosmological constant overtook matter and started to accelerate the expansion of the universe, see gure 4 for an illustration.

2.3 Ination

Ination is a phase of accelerated expansion of the universe which is assumed to have happened for a very short time after the Big Bang. The precise denition of ination is an epoch with accelerating scale factor, a¨ > 0. (2.11)

8 In order to have ination, one typically introduces a scalar eld with negative pressure, that drives the ination. We call this scalar eld the inaton ϕ. Let its Lagrangian be

1 L = − ϕ ϕ,µ − V (ϕ) (2.12) 2 ,µ with some potential V (ϕ). Then the inaton eld has the energy density

1 ρ = ϕ˙ 2 + V (ϕ) (2.13) ϕ 2 and the pressure 1 P = ϕ˙ 2 − V (ϕ). (2.14) ϕ 2

Note that for slowly varying ϕ, we have Pϕ ' −ρϕ, which gives wϕ ' −1, whence the inaton eld has negative pressure. Now we substitute the above equations into the Friedmann and Raychaudhuri equations, which gives:

  2 1 1 2 (2.15) H = 2 ϕ˙ + V (ϕ) , 3MPl 2 ϕ¨ + 3Hϕ = −V,ϕ, (2.16) where dV and is the reduced Planck mass, V,ϕ ≡ dϕ MPl

1 MPl ≡ √ . (2.17) 8πG One can show that these relations reduce to

2 1 (2.18) H ' 2 V (ϕ), 3MPl 3Hϕ ' −V,ϕ, (2.19) if the slow-roll conditions,

  1, (2.20) |η|  1, (2.21) are satised, where  and η are the slow-roll parameters, dened as:

M 2 V 2  ≡ Pl ,ϕ , (2.22) 2 V V η ≡ M 2 ,ϕϕ . (2.23) Pl V

9 dominating component w a(t) H(t) H−1 = (aH)−1 Λt −1 1 −Λt −1 Λ −1 a(t) ∼ e H(t) = Λ H = Λ e ∼ a radiation 1 1/2 1 −1 1/2 3 a(t) ∼ t H(t) = 2t H = 2t ∼ a matter 2/3 2 −1 3 1/3 1/2 0 a(t) ∼ t H(t) = 3t H = 2 t ∼ a Table 1: Evolution of the scale factor, the Hubble parameter and the Hubble horizon during dark energy, radiation and matter domination.

To see the connection between slow-roll conditions and ination, rewrite the condition for ination as

a¨ = H˙ + H2 > 0, (2.24) a or, equivalently, H˙ − < 1. (2.25) H2 Substituting the slow-roll equations gives

H˙ M 2 V 2 − ' Pl ,ϕ = . (2.26) H2 2 V

Hence, if the slow-roll conditions are satised (  1), then ination takes place (a¨ > 0). The duration of ination is controlled by the magnitude of η. The potential has to be at enough (which corresponds to a small η) so that  stays small. As an example, consider the inaton potential

1 V (ϕ) = m2ϕ2. (2.27) 2 The slow-roll parameters are 2M 2  = η = Pl . (2.28) ϕ2 Hence, ination occurs as long as 2 2 , and it breaks down near the minimum of the potential, ϕ  2MPl when 2 2 [8]. ϕ ∼ 2MPl

2.4 Evolution of scales

Later we will discuss the evolution of density perturbations in Fourier space, that is, the perturbations on a given comoving scale k−1. An important question is if the considered scale is larger or smaller than the Hubble horizon at the time. The (comoving) Hubble horizon is given by the inverse of the comoving Hubble parameter, H−1 = (aH)−1. Let us evaluate how this horizon develops in time during dierent periods of the universe. In order to do this, we use eq. (2.9) in order to nd the scale factor a(t). The results are listed in table 1. Note that during ination (Λ-domination) the comoving Hubble scale decreases, or in other words, the horizon shrinks. During this period, all scales leave the horizon. Later, when ination ends, the scales start to re-enter the horizon during the radiation- and matter- dominated era, small scales before large scales (see gure 5). We normalize the horizon so that today

10

Inflation

e l

a horizon

c

s

g

n

i

v

o

m o c comoving scale

time

Figure 5: A given comoving scale k−1 leaves the horizon during ination and re-enters later, during radiation or matter domination. (not to scale!)

−1 −1 Scale k [Mpc] k [Mpc ] aH zH 4286 (horizon today) 2.33 · 10−4 1 0 1000 10−3 5.44 · 10−2 17.4 100 0.01 5.44 · 10−4 1837 74.23 (equality scale) 0.0135 3 · 10−4 3332 10 0.1 4.04 · 10−5 24.7 · 103 1 1 4.04 · 10−6 24.7 · 104 0.1 10 4.04 · 10−7 24.7 · 105

1 Table 2: Dierent scales with their scale factor aH and redshift zH = − 1 of horizon entry. The aH scale factor at radiation-matter equality is given for a ΛCDM model and the Hubble horizon today is given for h = 0.7.

−1 2 (at a = 1) it is the radius of the whole visible universe, RH ' 3000h Mpc . Thus,

−1 1/2 H = RH a for a ≥ aeq. (2.29)

We then normalize the horizon before aeq to match the above relation:

R H−1 = H a for a < a . (2.30) 1/2 eq aeq

Table 2 shows some scales with their scale factor of horizon entry aH .

2 −1 −1 h ' 0.7 is the reduced Hubble constant, dened by H0 = 100h km s Mpc .

11 3 The perturbed universe: Newtonian treatment

3.1 Denitions

In cosmology, there are two dierent types of coordinates used, proper coordinates r and comoving coordinates x. These are related to each other by the scale factor a,

r = ax. (3.1)

Consequently, we need to dene the two dierent Nabla operators, ∂ and ∂ , which are ∇ ≡ ∂x ∇r ≡ ∂r related to each other by 1 ∇ = ∇. (3.2) r a In the same way, we can dene two dierent times, related to each other by the scale factor:

dt = adτ. (3.3)

We call t the proper time and τ the conformal time. In our notation, a dot denotes a partial derivative with respect to proper time t, a prime with respect to conformal time τ. We dene the Hubble parameter a˙ and the conformal Hubble parameter a0 . H ≡ a H ≡ a The absolute velocity u is dened as

dr d(ax) da dx u ≡ = = dτ x + = Hx + v = Hr + v = v + v, (3.4) dt adτ a dτ H where vH ≡ Hr = Hx is the Hubble velocity or Hubble ow, while the comoving (peculiar) velocity dx measures the velocity relative to the Hubble ow. v ≡ dτ The convective derivative in the (t, r)-system is given by

d ∂ = + u · ∇ , (3.5) dt ∂t r which follows simply from the chain rule. In the (τ, x)-system the convective derivative is given by

d ∂ = + v · ∇. (3.6) dτ ∂τ

Because proper time and conformal time are related by dt = adτ, the convective derivative transforms as d 1 d = . (3.7) dt a dτ

3.2 Perturbed elds

Now we are going to consider perturbations up to the second order. We start from the continuity equation, Poisson equation and Euler equation on an expanding, spatially at FRW background. We will consider a universe lled only with pressureless matter and no cosmological constant. This model

12 is called the Einstein-de Sitter model. The relevant elds are the matter density ρ, the gravitational potential φ and the absolute velocity u. Now we split the elds into a background part (denoted with a superscript (0)) and a perturbed part. The perturbed part itself can be further split into rst order perturbations, second order perturbations (superscripts (1), (2)) and so on:

ρ = ρ(0) + δρ = ρ(0) + ρ(1) + ρ(2) + ... = ρ(0)(1 + δ(1) + δ(2) + ...), (3.8) φ = φ(0) + δφ = φ(0) + φ(1) + φ(2) + ..., (3.9) (1) (2) u = vH + v = vH + v + v + ..., (3.10) where we have dened the n-th order density contrast δ(n),

ρ(n) δ(n) ≡ . (3.11) ρ(0)

Here, is considered as the background velocity, and the peculiar velocity dx vH = Hr = Hx v = dτ 3 is assumed to be small, |v|  |vH |. Furthermore, according to the Helmholtz theorem , v can be separated into a part with zero curl (longitudinal part vk) and a part with zero divergence (transverse part v⊥):

v = vk + v⊥, with ∇ × vk = 0 and ∇ · v⊥ = 0. (3.12)

Dene a scalar potential v and a vector potential Ω such that:

vk ≡ ∇v and v⊥ ≡ ∇ × Ω. (3.13)

3.3 Perturbed Newtonian equations

3.3.1 Continuity equation

The continuity equation in physical coordinates and proper time is

∂ ρ + ∇ (ρu) = 0. (3.14) ∂t r Using the Leibniz rule we obtain

∂ ρ + u · ∇ ρ + ρ∇ · u = 0. (3.15) ∂t r r

The rst two terms of this equation together form the convective derivative of ρ,

d ρ + ρ∇ · u = 0. (3.16) dt r

3The Helmholtz decomposition exists and is unique under the assumption that |v| → 0 as |x| → ∞, see appendix A for a proof.

13 Now we go from the (t, r)-system to the (τ, x)-system,

1 d 1 ρ + ρ∇ · u = 0. (3.17) a dτ a

Multiplying this equation by and using the denitions of the conformal convective derivative d and a dτ the absolute velocity u, we nd: ρ0 + 3Hρ + ∇ · (ρv) = 0. (3.18)

To the background order, this is ρ(0)0 + 3Hρ(0) = 0, (3.19) which is the well known mass continuity equation for p = 0 and has the solution ρ(0) ∼ a−3. To the linear order we nd the relation

ρ(1)0 + 3Hρ(1) + ρ(0)∇ · v(1) = 0. (3.20)

Now we use ρ(n)0 = (ρ(0)δ(n))0 = ρ(0)δ(n)0 − 3Hρ(0)δ(n) (3.21) with n = 1 to nd: δ(1)0 + ∇ · v(1) = 0. (3.22)

Note that only the scalar part of v survives, because we take its divergence. Hence, we can also write

δ(1)0 + ∆v(1) = 0, (3.23) where ∆ ≡ ∇2 is the Laplace operator in comoving coordinates. To the second order we nd:

ρ(2)0 + 3Hρ(2) + ∇ · (ρ(1)v(1)) + ρ(0)∇ · v(2) = 0. (3.24)

Again, we make use of eq. (3.21) with n = 2 to nd:

δ(2)0 + ∇ · (δ(1)v(1)) + ∇ · v(2) = 0. (3.25)

3.3.2 Poisson equation

The Poisson equation connects the gravitational potential with the matter density. In proper coordi- nates it is

∆rφ = 4πGρ. In comoving coordinates it becomes ∆φ = 4πGa2ρ. (3.26)

14 To the background order, we have

∆φ(0) = 4πGa2ρ(0), (3.27) from which we nd the background value for φ,

2 φ(0) = πGa2ρ(0)x2 + C(t), (3.28) 3 where C(t) is an arbitrary function of time. To the linear order, we have ∆φ(1) = 4πGa2ρ(0)δ(1), (3.29)

and to the second order, we nd

∆φ(2) = 4πGa2ρ(0)δ(2). (3.30)

3.3.3 Euler equation

The Euler equation tells us how the velocity eld changes in time given a gravitational potential. In proper coordinates it is ∂ u + u · ∇ u = −∇ φ. (3.31) ∂t r r Using the convective derivative we can express it in the shorter form

du = −∇ φ. (3.32) dt r

In the (τ, x)-system the equation becomes after multiplying with the scale factor

du = −∇φ. (3.33) dτ Using the denitions of the total velocity and the convective derivative we nd

H0x + v0 + Hv + v · ∇v = −∇φ. (3.34)

To the background order, we have, using the solution for φ(0),

4 H0 = − πGa2ρ(0), (3.35) 3 which is the Raychaudhuri equation for p = 0. The perturbed order is v0 + Hv + v · ∇v = −∇δφ. (3.36)

Note that 1 v · ∇v = ∇v2 − v × ω, (3.37) 2

15 where we have dened the vorticity ω ≡ ∇ × v, (3.38) so that eq. (3.36) reads 1 v0 + Hv + ∇v2 − v × ω = −∇δφ. (3.39) 2 Taking the curl of this equation, we nd the vorticity equation:

ω0 + Hω = ∇ × (v × ω). (3.40)

This dierential equation for the vorticity ω has an interesting solution: If ω = 0 everywhere initially, then the vorticity will stay zero at all times. However, the situation changes if we include a pressure gradient in the perturbed part of the Euler equation,

1 1 v0 + Hv + ∇v2 − v × ω = −∇δφ − ∇p. (3.41) 2 ρ

Then the vorticity equation becomes

1 ω0 + Hω = ∇ × (v × ω) + ∇ρ × ∇p. (3.42) ρ2

The second term on the r.h.s. is called the baroclinic contribution. It is zero in the case of a vanishing entropy gradient, ∇S = 0 [9]. Hence, if we consider a model with irrotational, isentropic initial conditions and adiabatic evolution, it follows that ω = 0 always. In the Einstein-de Sitter model, which we consider here, there is no entropy at all. In order to dene entropy, one needs at least two dierent types of uids. Hence, it is reasonable to assume

ω = 0, (3.43) which is equivalent to

v⊥ = 0. (3.44)

4 Hence, in Newtonian theory, the velocity perturbation is a pure potential ow , v = vk = ∇v. Furthermore, it follows that the vector potential Ω is a solution of the Laplace equation, since

0 = ω = ∇ × (∇ × Ω) = ∇(∇ · Ω) − ∆Ω = −∆Ω, where we set ∇ · Ω = 0 without any loss of generality. To the linear order, the Euler equation reads

v(1)0 + Hv(1) = −∇φ(1), (3.45)

4 Indeed, a constant non-zero v⊥ would also give ω = 0, but this would be in contradiction to the isotropy of the universe. Also, the Helmholtz decomposition is not unique if we allow constant parts in the vector.

16 and to the second order, we nd

v(2)0 + Hv(2) + v(1) · ∇v(1) = −∇φ(2). (3.46)

3.3.4 Summary

The important results of this section are the perturbed parts of the continuity, Poisson and Euler equation. In comoving coordinates and conformal time these are:

0. order 1. order 2. order C ρ(0)0 + 3Hρ(0) = 0 δ(1)0 + ∇ · v(1) = 0 δ(2)0 + ∇ · (δ(1)v(1)) + ∇ · v(2) = 0 P ∆φ(0) = 4πGa2ρ(0) ∆φ(1) = 4πGa2ρ(0)δ(1) ∆φ(2) = 4πGa2ρ(0)δ(2) E 0 4 2 (0) (1)0 (1) (1) (2)0 (2) (1) (1) (2) H = − 3 πGa ρ v + Hv = −∇φ v + Hv + v · ∇v = −∇φ Table 3: Continuity equation (C), Poisson equation (P) and Euler equation (E) in background, linear and quadratic order perturbation theory.

3.4 Solutions

3.4.1 First order solutions

In linear order the equations decouple into scalar and vector parts. The decoupled equations are:

scalar part vector part continuity equation (1)0 (1) - δ + ∇ · vk = 0 Poisson equation ∆φ(1) = 4πGa2ρ(0)δ(1) - Euler equation (1)0 (1) (1) (1)0 (1) vk + Hvk = −∇φ v⊥ + Hv⊥ = 0 Table 5: Decoupled rst order Newtonian equations.

The vector contribution to v(1) can be neglected because it decays ∼ a−1 (this is a special case of the above statement that the vorticity vanishes at all orders in Newtonian theory), so we can set (1) (1). We do a spatial Fourier transformation of the remaining equations. In momentum space v = vk these equations become:

(1)0 (1) (3.47) δk + ik · vk = 0, 6 −k2φ(1) = δ(1), (3.48) k τ 2 k 2 v(1)0 + v(1) = −ikφ(1), (3.49) k τ k k where we have used the Friedmann equation to replace the expression 4πGa2ρ(0):

4 8πG 6 = H2 = ρ(0)a2 ⇒ 4πGρ(0)a2 = . τ 2 3 τ 2

17 The solutions can be obtained as follows: First, solve the Poisson equation for (1): δk

1 δ(1) = − τ 2k2φ(1). (3.50) k 6 k The Euler equation can be written as

1 (av(1))0 = −ikφ(1), (3.51) a k k whence 1 Z τ v(1) = −ik dτ 0aφ(1). (3.52) k a k Now we insert the solutions for (1) and (1) into the the time derivative of the continuity equation, δk vk (1)00 (1)0 and nd the following equation for (1): δk + ik · vk = 0 φk τ φ(1)0 + φ(1)00 = 0. (3.53) k 6 k

This is a second order dierential equation for (1) and has two solutions. Whenever this is the case, φk we call the dominating solution the growing mode and the subdominant solution the decaying mode. Here, the decaying solution is (1) −5 and the growing solution is a constant, φk ∼ τ

(1) (1) (3.54) φk = φk (k).

In summary, the growing rst order solutions are:

(1) (1) (3.55) φk = φk (k), 1 2 δ(1) = − τ 2k2φ(1) = − ak2R2 φ(1)(k), (3.56) k 6 k 3 H k i 2 v(1) = − kτφ(1) = − ika1/2R φ(1)(k). (3.57) k 3 k 3 H k

Note that, in order to convert τ to a, we use the denition of the Hubble parameter:

2 2 2 t0 −3/2 −1/2 = H = aH = a = a = aH0a = H0a , τ 3t 3t0 t whence 1/2 −1 1/2 c −1 1/2 −1 τ = 2a H0 = 2a c = 2a RH c , H0 c where = RH ' 4300Mpc is the Hubble radius. Here, we use c = 1-units, so that the connection H0 between τ and a is simply 1/2 τ = 2a RH . (3.58)

18 The decaying solutions for (1) and (1) can be obtained using the decaying solution for (1), δk vk φk

(1) −5 (3.59) φk ∼ τ , (1) −3 (3.60) δk ∼ τ , (1) −4 (3.61) vk ∼ τ .

3.4.2 Initial conditions

Now we want to answer the question how big φk is, since all the other quantities depend on it (we drop the superscripts ...(1) for the moment). For this we have to go deeper into the theory of the primordial power spectrum. Note that this is not part of Newtonian theory. Ination creates curvature perturbations. Denote the curvature perturbation eld by ζ(x, τ), and its Fourier transform by ζk. The precise denition will be given in subsection 4.2.4. An important physical quantity is its power spectrum. It is dened by

k3 P (k) ≡ |ζ |2. (3.62) ζ 2π2 k Measurements of WMAP give a nearly scale-invariant (Harrison-Zel'dovich) power spectrum for the primordial curvature perturbation [10],

2 −5 Pζ (k) = A ,A ' 5 · 10 . (3.63)

Bardeen showed in [11] that in the matter dominated era of the universe the Bardeen potential Φk (we will give the denition later) is related to ζk via

3 Φ = − ζ , (3.64) k 5 k

As we will show later, the Bardeen potential Φk can be identied with the gravitational potential perturbation φk. Hence, φk has the following power spectrum:

k3 9 k3 9 9 P (k) = |φ |2 = |ζ |2 = P (k) = A2. (3.65) φ 2π2 k 25 2π2 k 25 ζ 25

Thus, also the spectrum for φk is scale-invariant. However, for the spectrum of δk we nd a scale- dependence:

k3 k3 4 4 4 P (k) = |δ |2 = a2k4R4 |φ |2 = a2k4R4 P (k) = a2k4R4 A2 ∼ k4, (3.66) δ 2π2 k 2π2 9 H k 9 H φ 25 H or equivalently, 8 P (k) = π2R4 A2a2k ∼ k. (3.67) δ 25 H This means that there is more power (and therefore more structure) on smaller scales. Measurements −1 show that the k-dependence of Pδ(k) is correct on large scales, but at a scale of about keq ' 0.01 Mpc ,

19 Figure 6: The matter power spectrum today, measured using various techniques [12]. The bend −1 at k = keq ' 0.01 Mpc is a result of the transition between radiation domination and matter domination. which is the scale that enters the horizon at the time of radiation-matter equality, the power spectrum has a bend and decreases roughly as k−3 for higher k (see gure 6). This bend comes from the fact that the universe was not always matter-dominated, but there was a radiation-dominated era (which is not included so far in our model). In order to explain the bend, we have to nd out how density perturbations grow during radiation domination.

3.4.3 Transfer function

As we discussed above, it is important that in the history of the universe there was a radiation- dominated era, where the pressure was not zero. Therefore we consider now matter density pertur- bations during radiation domination. However, we are going to neglect perturbations in the radiation density, which turns out to be a good approximation. If we consider a universe that contains both matter and radiation, we have to modify the Euler equation, containing both a gravitational potential gradient and a pressure gradient. Then the rst order Newtonian equations are:

δ(1)0 + ∇ · v(1) = 0, (3.68) (1) 2 (1) ∆φ = 4πGa ρ¯mδ , (3.69) p(1) v(1)0 + Hv(1) = −∇(φ(1) + ), (3.70) ρ¯

20 where ρ¯ =ρ ¯m +ρ ¯γ is the total background energy density. Now we take the time derivative of eq. (3.68) and subtract the divergence of eq. (3.70), using eq. (3.69) to replace ∆φ(1). This gives the following relation: (1)00 (1)0 2 (1) 2 (1) (3.71) δ + Hδ = 4πGρ¯ma δ + ∆csδ , where 2 p(1) is the square of the speed of sound. In Fourier space this equation reads: cs = w = ρ¯

(1)00 (1)0 2 (1) 2 2 (1) (3.72) δk + Hδk = 4πGρ¯ma δk − k csδk .

This equation is known as the growth equation or Jeans equation for (1) [13]. The l.h.s. represents the δk time change of (1), including a friction term (1)0, which is also called Hubble damping or Hubble δk Hδk friction (the expansion of the universe slows down the growth of density perturbations). On the r.h.s. there are two competitive sources: , which supports the growth of density perturbations, and pressure, which prevents it. It is convenient to introduce the comoving Jeans wavenumber,

 2 1/2 4πGρ¯ma (3.73) kJ ≡ 2 , cs and the corresponding comoving Jeans length or Jeans scale,

 2 −1/2 −1 4πGρ¯ma (3.74) λJ ≡ kJ = 2 . cs Then the Jeans equation can be written as

(1)00 (1)0 2 2 2 (1) (3.75) δk + Hδk = (kJ − k )csδk .

Now we have two cases:

1. For scales much smaller than the Jeans length, k  kJ , we can approximate the Jeans equation by (1)00 (1)0 2 2 (1) (3.76) δk + Hδk = −k csδk .

The solutions are oscillating with the frequency kcs, and the oscillations are damped by the Hubble friction term, whence the amplitude of the oscillations decays with time. There is no structure growth on sub-Jeans scales.

2. For scales much larger than the Jeans length, k  kJ , we can approximate the Jeans equation by (1)00 (1)0 2 (1) (3.77) δk + Hδk = 4πGρ¯ma δk . During matter domination, this gives the growing solution (1) , which we already know. δk ∼ a During radiation domination, we have

ρ¯m  ρ¯ ' ρ¯γ (3.78)

21 and hence 2 3 2 ρ¯m 4πGa ρ¯m = H  1, (3.79) 2 ρ¯γ so that we can neglect the source term on the right hand side. Then the Jeans equation becomes

(1)00 (1)0 (3.80) δk + Hδk = 0,

which has the general solution (1) (3.81) δk = C1 + C2 ln a. Hence, during radiation domination density perturbations grow at most logarithmically on super- Jeans scales.

Now we estimate the Jeans scale. During matter domination we have λJ = 0 since cs = 0, so that q 1 perturbations grow on all scales. During radiation domination the speed of sound is cs = and √ 3 p 2 therefore kJ = 3 4πGρ¯ma . As we approach the time of radiation-matter equality, we have ρ¯m ≈ ρ¯, whence . Thus, the Jeans scale grows to the size of the horizon at radiation-matter ρ¯γ ≈ 2 kJ ≈ H equality and reduces to zero when matter dominates. Consequently, it makes a dierence if a scale enters the horizon during radiation domination or during matter domination because perturbations on scales that enter during radiation domination experience a suppression in growth. Now we want to quantify this eect. First consider a density perturbation on a scale −1 −1, which enters the k > keq horizon during matter domination. At a later time af during matter domination it has grown by a factor: δ(1)(a ) H2 k2 k f = enter = . (3.82) (1) H2 H2 δk (aenter) f f Now consider a perturbation on a scale −1 −1. This scale enters the horizon during radiation k < keq domination. But then, between the time it enters and the time matter takes over nothing happens to the perturbation. Thus,

δ(1)(a ) δ(1)(a ) H2 k2 k2 k2 k f = k f = eq = eq = eq . (3.83) (1) (1) H2 H2 k2 H2 δk (aenter) δk (aeq) f f f Hence, the growth on this scale is suppressed by a factor

k2 T (k) ≡ eq (3.84) k2 compared to a scale that enters during matter domination. In the last equation we introduced the transfer function , which gives the decit in growth that perturbations on scales smaller than −1 T (k) keq experience. It should have the following properties:

2 T (k) = 1 for k  keq, (3.85) k4 T 2(k) ∼ eq for k  k . (3.86) k4 eq

22 1 simple transfer function BBKS transfer function 0.9

0.8

0.7

0.6

0.5 T(k)

0.4

0.3

0.2

0.1

0 1e-04 0.001 0.01 0.1 1 k [Mpc-1]

−1 2  k4  Figure 7: Two dierent transfer functions. The simple one, T (k) = 1 + 4 , cuts away too much keq power on small scales, because it neglects the growth of the CDM density contrast during radiation domination on these scales.

A simple choice is 1 T 2(k) = . (3.87) k4 1 + 4 keq

If we want to make the transfer function more accurate, we have to account for cold dark matter (CDM) perturbations. Since CDM does not interact with radiation, perturbations do not feel the radiation pressure and grow independently of the Jeans scale (see appendix D). Hence, our simplied transfer function suppresses too much power on small scales. The correction is roughly logarithmic, so that the transfer function should go as:

2 T (k) = 1 for k  keq, (3.88) k4 2 eq 2 k for (3.89) T (k) ∼ 4 ln ( ) k  keq. k keq

The most prominent transfer function that includes this correction and gives power spectra in good agreement with measurements is the BBKS transfer function, introduced in 1986 by Bardeen, Bond, Kaiser and Szalay [14],

ln(1 + 2.34q) T (k) = [1 + 3.89q + (16.1q)2 + (5.46q)3 + (6.71q)4]−1/4, (3.90) BBKS 2.34q

23 where k and the shape parameter is given by q ≡ ΓhMpc−1 Γ

√ Ωb Γ ≡ Ω0h exp(−Ωb − 2h ), (3.91) Ω0 where Ω0 is the relative energy density of the universe today and Ωb is the relative baryon density today. Henceforth, we will use the BBKS transfer function with the parameters Ω0 = 1, Ωb = 0.05 and h = 0.7. We show a plot of the two dierent transfer functions in gure 7. The transfer function connects the real power spectrum and the power spectrum from the Einstein- de Sitter model. Hence, we make the following transformation:

P(k) → T 2(k)P(k), (3.92)

The transfer function explains the bend in matter power spectrum (see gure 11). Furthermore, now we can answer the question what the initial value for (1) is. Including the transfer function, eq. (3.65) φk becomes k3 9 P (k) = |φ(1)|2 = A2T 2(k). (3.93) φ 2π2 k 25 From this we nd the following initial value for (1): φk

3 |φ(1)| = A(2π2)1/2k−3/2T (k). (3.94) k 5

3.4.4 Second order solutions

The second order equations do not decouple a priori; we have to neglect the vector parts of the rst order quantities, which can be done since (1) decays. The decoupled second order equations are: v⊥

scalar part vector part continuity equation (2)0 (2) (1) (1) - δ + ∇ · vk = −∇ · (δ vk ) Poisson equation ∆φ(2) − 4πGa2ρ(0)δ(2) = 0 - Euler equation (2)0 (2) (2) (1) (1) (2)0 (2) vk + Hvk + ∇φ = −vk · ∇vk v⊥ + Hv⊥ = 0 Table 6: Decoupled second order equations.

The vector part of the Euler equation gives (2) −1 (again, this is a special case of the general v⊥ ∼ a result that the vorticity vanishes at all orders in Newtonian theory, as we showed earlier), so we can neglect the vector contribution to (2) and set (2) (2). Now consider the scalar equations: The v v = vk sources for the second order perturbations are products of rst order perturbations, which we have written on the r.h.s. of each equation. There are hence two solutions. The homogeneous solutions are obtained by ignoring the source terms, which gives the same time dependence as in rst order perturbation theory. However, the specic solutions grow faster due to the source terms, so that we 2 can neglect the homogeneous solutions. Note that the term (1) (1) can also be written as v(1) . v · ∇v ∇ 2 Now we need to be careful when transforming to momentum space, because products in x-space become convolutions in k-space.

24 2 • In momentum space the expression v(1) becomes:

F[v(1) · v(1)](k) 1 = (v ? v)(k) (2π)3 1 Z = d3k0v(k0) · v(k − k0) (2π)3 1 1 Z = − τ 3 d3k0k0φ(k0) · (k − k0)φ(|k − k0|) (2π)3 9 Z ∞ Z π 1 1 3 0 0 02 0 0 0 02 0 p 2 02 0 0 = − 2 τ dk dϑ k sin ϑ (kk cos ϑ − k )φ(k )φ( k + k − 2kk cos ϑ ). (2π) 9 0 0

• The expression ∇ · (δ(1)v(1)) becomes:

F[∇ · δ(1)v(1)](k) 1 =ik · (δ(1) ? v(1))(k) (2π)3 1 Z =ik · d3k0δ(1)(k0)v(k − k0) (2π)3 1 i Z =ik · τ 3 d3k0k02φ(k0)φ(|k − k0|)(k − k0) (2π)3 18 1 1 Z = − τ 3 d3k0k02φ(k0)φ(|k − k0|)k · (k − k0) (2π)3 18 Z ∞ Z π 1 1 3 0 0 04 0 2 0 0 0 p 2 02 0 0 = − 2 τ dk dϑ k sin ϑ (k − kk cos ϑ )φ(k )φ( k + k − 2kk cos ϑ ), (2π) 18 0 0

0 sin ϑ0 cos ϕ0 where we set   and 0 0  0 0 , so that 0 0 0. k = k 0 k = k sin ϑ sin ϕ  k · k = kk cos ϑ 1 cos ϑ0

The second order evolution equations in momentum space are:

(2)0 (2) (1) (1) (3.95) δk − ik · vk = −F[∇ · (δ v )], 6 k2φ(2) + δ(2) = 0, (3.96) k τ 2 k 2 1 v(2)0 + v(2) + ikφ(2) = −ik F[v(1) · v(1)]. (3.97) k τ k k 2

Note that δ(1)v(1) ∼ τ 3 and v(1) · v(1) ∼ τ 2 in leading order. There are other solutions as well, which can be obtained by combining decaying and growing modes of δ(1) and v(1). However, these solutions are all decaying and can be neglected. Hence, it is reasonable to make the following ansatz by adjusting

25 the time dependence of the second order perturbations to the time dependence of the source terms:

(2) 4 (3.98) δk = Ckτ , k v(2) = D i τ 3, (3.99) k k k (2) 2 (3.100) φk = Ekτ .

With this ansatz, we nd the following system of linear equations for the coecients Ck, Dk, Ek:

(1) (1) −3 4Ck − Dkk = −F[∇ · δ v ](k)τ , (3.101) 2 Ekk + 6Ck = 0, (3.102) 1 5D + E k = − kF[v(1) · v(1)](k)τ −2. (3.103) k k 2 The solutions are obtained by rst solving the convolution integrals and then solving the system of linear equations for the coecients. This is done numerically for three xed scales that we are interested in: the galaxy scale (k−1 = 1 Mpc), the cluster scale (k−1 = 10 Mpc), and the supercluster scale (k−1 = 100 Mpc). Figures 8, 9 and 10 show the plots of the rst and second order perturbations (1) √ 3/2 (1) on these scales. Note that we plot the dimensionless quantities ∆ ≡ P = √k |δ | etc. δ δ 2π2 k

100

1

0.01 )

-1 1e-04

1e-06 (a, k=1 Mpc ∆

1e-08

Φ 1 v 1e-10 δ1 Φ1 2 v δ2 2 1e-12 1e-04 0.001 0.01 0.1 1 a

Figure 8: Evolution of rst and second order perturbations for the scale k−1 = 1 Mpc.

26 100

1

0.01 ) -1 1e-04

1e-06 (a, k=0.1Mpc ∆

1e-08

Φ 1 v 1e-10 δ1 Φ1 2 v δ2 2 1e-12 1e-04 0.001 0.01 0.1 1 a

Figure 9: Evolution of rst and second order perturbations for the scale k−1 = 10 Mpc.

1

0.01

1e-04

) 1e-06 -1

1e-08 (a, k=0.01Mpc

∆ 1e-10

1e-12 Φ 1 v1 1e-14 δ Φ1 2 v δ2 2 1e-16 1e-04 0.001 0.01 0.1 1 a

Figure 10: Evolution of rst and second order perturbations for the scale k−1 = 100 Mpc.

27 scale 1 Mpc 10 Mpc 100 Mpc

aNL1 0.005 0.092 9 aNL2 0.12 0.75 30

Table 7: aNL1 and aNL2 for dierent scales.

3.4.5 Discussion

For early times and large scales linear perturbation theory is valid, since all second order perturbations are substantially smaller than all rst order perturbations. Now a diculty arises to dene the time when a scale goes non-linear. Strictly speaking, linear perturbation theory becomes invalid when the largest second order contribution becomes as important as the smallest rst order contribution. Here this is the crossing of (2) and (1). We call the scale factor at this crossing , δk φk aNL1

(2) (1) (3.104) δk (aNL1) ' φk (aNL1).

However, the perturbations series (1) (2) still makes sense, since (2) is much smaller than δk = δk + δk δk (1) at this time. According to the spherical collapse model, collapse happens much later, roughly δk when the density contrast reaches unity [8, 13], which is a commonly accepted indicator for the begin of nonlinearity. A density contrast of 1 means that the density perturbations become as large as the background density. In other words, there can be regions which become twice as dense as the background. We call the scale factor when this happens aNL2,

(1) (3.105) ∆δ (aNL2) ' 1.

The transition between the the linear system and the nonlinear system happens certainly between aNL1 and aNL2 (see table 7). Shortly after aNL2, the system has viralized into a gravitational bound object, where the angular momentum prevents further collapse [8]. This has already happened for galaxies and clusters of galaxies, while perturbations on scales & 10 Mpc still behave linear today. The fact that small scales collapse before large scales leads to the bottom-up picture of structure formation: small scale structures form before large scale structures. Figure 11 shows the evaluated matter power spectrum (1) 2 today. It increases for Pδ(k) = |δk | ∼ k −3 2 k < keq and decreases ∼ k ln (k) for k > keq due to the implemented transfer function and shows good correspondence with measurements, as shown in gure 6. However, nonlinear eects become −1 important as Pδ(k) reaches unity, which happens at about k = 0.078 Mpc today. Figure 12 shows the power spectrum for the absolute value of the peculiar velocity (1) 2 today. It decreases Pv(k) = |vk | −1 −5 2 ∼ k for k < keq and ∼ k ln (k) for k > keq.

28 100

1

0.01

1e-04 (a=1, k) 2 δ

∆ 1e-06

1e-08

1e-10

1e-12 1e-04 0.001 0.01 0.1 1 k [Mpc-1]

100000

10000 ] 3 [Mpc δ P

1000

100 1e-04 0.001 0.01 0.1 1 k [Mpc-1]

Figure 11: The matter power spectrum today. The top gure shows Pδ(k) and the bottom gure shows . Linear theory fails when (vertical line). Pδ(k) Pδ(k) & 1

29 1e-04

1e-05

1e-06

1e-07 (a=1, k) 2 v

∆ 1e-08

1e-09

1e-10

1e-11 1e-04 0.001 0.01 0.1 1 k [Mpc-1]

10000

1000

100

10 ]

3 1 [Mpc v

P 0.1

0.01

0.001

1e-04

1e-05 1e-04 0.001 0.01 0.1 1 k [Mpc-1]

Figure 12: The power spectrum of |v| today. The top gure shows Pv(k) and the bottom gure shows Pv(k).

30 3.4.6 Solutions during Λ-domination

As we approach a = 1, the cosmological constant becomes more and more important, as matter scales ∼ a−3. During Λ-domination the Friedmann and Raychaudhuri equations become

8πG H2 = a2ρ , (3.106) 3 Λ 8πG H0 = a2ρ . (3.107) 3 Λ However, if we neglect dark energy perturbations, there are no variations in the rst order equations. Hence, the form of the Jeans equation remains unchanged,

(1)00 (1)0 2 (1) (3.108) δk + Hδk = 4πGρ¯ma δk .

Note that 2 3 2 ρ¯m 4πGρ¯ma = H  1, (3.109) 2 ρΛ so that the source term of the Jeans equation becomes negligible. Then we have two solutions, the growing solution (1) const. (3.110) δk = and the decaying solution (1) −2 (3.111) δk ∼ a .

Hence, as we approach Λ-domination, the density perturbations freeze out. A more general solution in the intermediate regime, when dark energy and matter density are equally important, will be discussed later, in section 5.3.

3.5 Newtonian cosmological simulations

Cosmological N-body simulations use Newton's equation of motion to simulate the behaviour of a xed number of particles under the inuence of their reciprocal gravitational interaction. These simulations are usually run in a box with periodic boundary conditions. In the (t, r)-system the equation of motion can be written as d2r (t) i = −∇ φ , (3.112) dt2 r sim where ri(t) is the trajectory of the i-th particle. Transforming to the (τ,x)-system, we nd

d d 1 (ax (τ)) = ∇φ . (3.113) adτ adτ i a sim The l.h.s. is

d d d  dx (τ) 1  dx (τ) dx2(τ) (ax (τ)) = Hx (τ) + i = H0x (τ) + H i + i , adτ adτ i adτ i dτ a i dτ dτ 2

31 and on the r.h.s. the potential can be split into a background part and a perturbed part,

1 1 ∇φ = ∇(φ¯ + δφ ). a sim a sim sim Subtracting the background (this is the Raychaudhuri equation), we nd

d2x (τ) dx (τ) i + H i = −∇δφ . (3.114) dτ 2 dτ sim

From this equation it can be seen that all freely falling observers (for which δφsim = 0) will become resting observers after suciently long time, since dxi(τ) 1 . The gravitational potential perturbation dτ ∼ a is obtained using the Poisson equation,

2 ∆δφsim = 4πGa δρsim, (3.115) which is solved in Fourier space, where the matter density perturbation is calculated by counting particles in cells, −3 X δρsim(x, τ) = a miδD(x − xi(τ)). (3.116) i

4 The perturbed universe: relativistic treatment

4.1 The perturbed metric

In relativistic cosmological perturbation theory we consider small perturbations on the homogenous and isotropic FRW metric tensor and on the energy-momentum tensor and calculate how these per- turbations develop in time, using the energy-momentum conservation equation and the Einstein eld equations. For a detailed reference, see [15]. Our conventions are the following: Greek indices will range from 0 to 3 and Latin indices from 1 to 3. ∇µ as well as an index ; µ denote a covariant derivative with respect to the perturbed metric. We are going to drop the superscripts ...(1) here, since we only consider linear perturbations. Background quantities are denoted with a bar. The FRW metric in a at universe, using comoving coordinates and conformal time, is

! 2 −1 0 g¯µν ≡ a . (4.1) 0 δij

Now consider a small perturbation of the background metric: gµν =g ¯µν + δgµν . We dene the metric perturbation ! 2 −2φ wi δgµν ≡ a . (4.2) wi −2eδij + 2hij wi and hij can be further decomposed into scalar, vector and tensor parts,

⊥ (4.3) wi = w;i + wi

32 type elds constraints degrees of freedom scalar perturbations φ,ψ,w,h - 4 vector perturbations ⊥ i ⊥ i 4 wi , hi ∇ wi = ∇ hi = 0 tensor perturbations T i T T i , T T 2 hij ∇ hij = (h )i = 0 hij = hji Table 8: Degrees of freedom in linear perturbation theory. and

T (4.4) hij = Dijh + h(i;j) + hij, where Dij is the symmetric traceless double-gradient operator,

1 D ≡ ∇ ∇ − δ ∆. (4.5) ij i j 3 ij

Note that ⊥ and are pure vector parts with zero divergence ( i ⊥ i ) and T is wi hi ∇ wi = ∇ hi = 0 hij transverse ( i T ), traceless ( T i ) and symmetric ( T T ). Thus, in rst order the ∇ hij = 0 (h )i = 0 hij = hji metric perturbations decouple:

S V T (4.6) δgµν = δgµν + δgµν + δgµν ! ! ! −2φ w 0 w⊥ 0 0 = a2 ;i + a2 i + a2 , (4.7) ⊥ T w;i −2ψδij + 2h;ij wi 2h(i;j) 0 2hij where we have introduced a new variable5,

1 ψ ≡ e + ∆h. (4.8) 3 Now that we have decomposed the metric perturbation, it is instructive to count the degrees of freedom, see table 8. The result is that we have altogether 10 degrees of freedom in the metric perturbation. However, four of them correspond to the freedom of coordinate choice, so that we can put further restrictions on the metric, as will be discussed later. The decomposition into scalar, vector and tensor parts is important for the physical interpretation. The scalar perturbations are the most important, because they couple to density perturbations and give rise to structure formation. The vector per- turbations are not very interesting in linear order, because they couple to rotations only, which are decaying in an expanding universe due to angular momentum conservation. Tensor perturbations give rise to gravitational waves.

4.2 Gauge transformations

There is no unique mapping between points in the background spacetime and points in the perturbed spacetime. Indeed, there are innitely many mappings, all close to each other. Consequently, for

5Some authors introduce this quantity as the curvature perturbation. However, we will later introduce the curvature perturbation according to the denition of Bardeen [11].

33 ~ ̃ ̂ P P P̄. .. ^

α x̂

α x̃ α ̄x

Figure 13: Illustration of two dierent gauge choices.

a given coordinate system of the background, there exist innitely many coordinate systems of the perturbed spacetime. A coordinate transformation between these coordinate systems is called a gauge transformation. Consider a point P¯ in the background spacetime, whose coordinates are {xα}. Now consider two dierent coordinate systems in the perturbed spacetime, and denote them by {x˜α} and {xˆα}. In the x˜α-coordinates, the background point P¯ corresponds to a point P˜(˜x). In the xˆα- coordinates, this point corresponds to another point Pˆ(ˆx), see gure 13 for an illustration. We have by construction xα(P¯) =x ˜α(P˜) =x ˆα(Pˆ), (4.9) because all coordinates refer to the same point in the manifold. Now one can ask the question: what are the xˆα-coordinates of the point P˜? These clearly dier from the x˜α-coordinates and we denote this dierence by ξα, which is rst-order small due to the fact that all coordinate systems are close to each other: x˜α(P˜) =x ˆα(P˜) + ξα, (4.10)

x˜α(Pˆ) =x ˆα(Pˆ) + ξα. (4.11)

Note that it does not matter, in which coordinate system we give ξα, because the dierence is second order. The above equations represent a coordinate transformation, which in cosmological perturbation theory is called gauge transformation. They are equivalent to:

xˆα(P˜) =x ˆα(Pˆ) − ξα, (4.12)

x˜α(P˜) =x ˜α(Pˆ) − ξα. (4.13)

34 4.2.1 Gauge transformations of scalars, vectors and tensors

Using these transformations, we can determine how various geometric objects (scalars, vectors and tensors) transform under a gauge transformation.

• First, let us determine the transformation of a scalar. Using Taylor expansion, we nd

˜ ˆ ∂s α ˜ α ˆ (1) ˆ ∂s¯ α (2) ˆ 0 0 (4.14) s(P ) = s(P ) + α (ˆx (P ) − xˆ (P )) = s(P ) − α ξ = s(P ) − s¯ ξ . ∂xˆ | {z } ∂x −ξα

Here we have used that (1) the dierence between ∂s and ∂s¯ is rst order small and can be ∂xˆα ∂xα neglected when multiplied by ξα and (2) the background value s¯ only depends on conformal time τ, because the background spacetime is isotropic. The scalar perturbation in a given gauge is dened as δs˜ ≡ s(P˜) − s¯(P¯) (or δsˆ ≡ s(Pˆ) − s¯(P¯)). The transformation law is

δs˜ = s(P˜) − s¯(P¯) = s(Pˆ) − s(P¯) − s¯0ξ0 = δsˆ − s¯0ξ0. (4.15)

Hence, any scalar that is constant in the background (s¯0 = 0) is gauge-invariant.

µ • For a vector wµˆ(Pˆ) we have (µˆ refers to the coordinates xˆ ):

∂w¯ w (P˜) = w (Pˆ) − µ ξα (4.16) µˆ µˆ ∂xα and therefore

∂xˆσ ∂(˜xσ − ξσ) ∂w¯ w (P˜) = w (P˜) = (w (Pˆ) − σ ξα) µ˜ ∂x˜µ σˆ ∂x˜µ σˆ ∂xα ∂w¯ = (δσ − ξσ )(w (Pˆ) − σ ξα) µ ,µ σˆ ∂xα ˆ α σ (4.17) = wµˆ(P ) − w¯µ,αξ − w¯σξ,µ.

For a vector perturbation we nd the transformation

˜ δwµ = wµ˜(P˜) − w¯µ(P¯) ˆ α σ (4.18) = δwµ − w¯µ,αξ − w¯σξ,µ.

Obviously, any vector that vanishes in the background is gauge-invariant.

• For a (0,2)-tensor Bµˆνˆ(Pˆ) we have:

α α Bµˆνˆ(P˜) = Bµˆνˆ(Pˆ) − Bµˆνˆ,αˆ ξ = Bµˆνˆ(Pˆ) − B¯µν,αξ (4.19)

35 and therefore

∂xˆρ ∂xˆσ B (P˜) = B (P˜) µ˜ν˜ ∂x˜µ ∂x˜ν ρˆσˆ ∂(˜xρ − ξρ) ∂(˜xσ − ξσ) = (B (Pˆ) − B¯ ξα) ∂x˜µ ∂x˜ν ρˆσˆ ρσ,α ρ ρ σ σ ˆ ¯ α = (δµ − ξ,µ)(δν − ξ,ν )(Bρˆσˆ (P ) − Bρσ,αξ ) ˆ ¯ α ρ ¯ σ ¯ (4.20) = Bµˆνˆ(P ) − Bµν,αξ − ξ,µBρν − ξ,ν Bµσ.

For a tensor perturbation we nd

˜ δBµν = Bµ˜ν˜(P˜) − B¯µν (P¯) ˆ ¯ α ρ ¯ σ ¯ (4.21) = δBµν − Bµν,αξ − ξ,µBρν − ξ,ν Bµσ.

Again, any tensor that vanishes on the background is gauge-invariant.

4.2.2 Gauge transformation of metric perturbations

Now we can evaluate how the metric perturbation transforms. Using that the background metric is symmetric and homogeneous, eq. (4.21) gives:

˜ ˆ 0 0 (4.22) δgµν = δgµν − g¯µν ξ − 2ξ(µ,ν).

• The (0, 0)-component of eq. (4.22) reads, using the metric tensor from eq. (4.1) and the perturbed part of the metric from eq. (4.2):

−2a2φ˜ = −2a2φˆ + 2aa0ξ0 + 2a2ξ00

⇒ φ˜ = φˆ − Hξ0 − ξ00. (4.23)

• From the (0, i)-component of eq. (4.22) we nd the transformation of wi:

2 2 −a w˜i = −a wˆi − 2ξ(0,i) 2 2 0 2 0 = −a wˆi − a ξ,i + a ξi

0 0 (4.24) ⇒ w˜i =w ˆi + ξ,i − ξi.

This can be further decomposed using ⊥, ξi = ξ,i + ξi

w˜ =w ˆ + ξ0 − ξ0, (4.25) ⊥ ⊥ ⊥0 (4.26) w˜i =w ˆi − ξi .

36 • The (i, j)-component of eq. (4.22) is:

2 2˜ 2 2ˆ 0 0 − 2a eδ˜ ij + 2a hij = −2a eδˆ ij + 2a hij − 2aa δijξ − 2ξ(i,j). (4.27)

Note that ξ(i,j) can be split into a traceless part and a diagonal part:

1 1 ξ = ξ δ + ξ − ξ δ . (4.28) (i,j) 3 k,k ij (i,j) 3 k,k ij | {z } traceless

The trace of eq. (4.27) gives the transformation of e:

1 e˜ =e ˆ + Hξ0 + ξk . (4.29) 3 ,k

If we take the traceless part of eq. (4.27), we nd the transformation of hij:

1 h˜ = hˆ − ξ(i + δ ξk . (4.30) ij ij ,j) 3 ij ,k This can be further decomposed,

h˜ = hˆ − ξ, (4.31) ˜ ˆ ⊥ (4.32) hi = hi − ξi , ˜T ˆT (4.33) hij = hij.

• From eq. (4.31) and eq. (4.29) we nd the transformation law for ψ,

ψ˜ = ψˆ + Hξ0. (4.34)

4.2.3 Gauge transformations of velocity and density perturbations

Next, consider the transformation of the peculiar velocity v. We have:

dx˜i d(ˆxi + ξi) d(ˆxi + ξi) d(ˆτ + ξ0)−1 v˜i = = = dτ˜ d(ˆτ + ξ0) dτˆ dτˆ = (ˆvi + ξi0)(1 + ξ00)−1 = (ˆvi + ξi0)(1 − ξ00) =v ˆi + ξi0, so that the velocity potential transforms as

v˜ =v ˆ + ξ0, (4.35) and the vector part transforms as ⊥ ⊥ ⊥0 (4.36) v˜i =v ˆi + ξi .

37 For the density contrast δ we nd the transformation:

δρ˜ δρˆ − ρ¯0ξ0 ρ¯0 δ˜ = = = δˆ − ξ0 = δˆ + 3Hξ0. (4.37) ρ¯ ρ¯ ρ¯

This means that v and δ depend on the gauge choice; they are dierent in each gauge. Hence, they cannot be unique physical observables. However, it possible to construct gauge-invariant quantities, which do not depend on the gauge choice, as we will show in the next subsection.

4.2.4 Gauge-invariant quantities

As mentioned before, unique physical quantities need to be gauge-invariant. A good example for this is the theory of electromagnetism. Consider the free Lagrangian,

1 L = − F µν F , (4.38) EM 4 µν where Fµν ≡ 2A(µ,ν) is the electromagnetic tensor and Aµ is the electromagnetic 4-potential. Note that the Lagrangian is invariant under the gauge transformation

Aµ → Aµ + χ,µ. (4.39)

However, we can construct two gauge-invariant quantities out of Aµ, the electric eld

E ≡ A˙ + ∇A0 (4.40) and the magnetic eld B ≡ ∇ × A. (4.41)

We can learn from this example that the question of gauge only arises in the theory. Once we evaluate measurable quantities, the gauge should become irrelevant. However, here we ignored the role of the observer. For example, if we consider a charge at a xed point, a resting observer would measure no magnetic eld, while a moving observer would. This shows that also in electrodynamics the role of the observer plays a crucial role. Now let us come back to cosmological perturbation theory. We can construct gauge-invariant combinations out of the perturbed quantities. The most famous ones are the Bardeen potentials [11],

1 Φ ≡ φ + [(w − h0)a]0 , (4.42) a Ψ ≡ ψ − H(w − h0). (4.43)

Another important gauge-invariant quantity is the curvature perturbation, which we dene according to Bardeen [11], 1 ζ ≡ δ + ψ. (4.44) 3 We introduced this quantity already in subsection 3.4.2, where we used its power spectrum to nd the

38 φ = 0 ψ = 0 δ = 0 w = 0 h = 0 v = 0 ∗ ∗ φ = 0 - / / S ∗∗ ∗∗ ∗∗ ∗ ∗ ψ = 0 / - / UC SE ∗ δ = 0 / / - UD ∗ ∗ ∗ w = 0 S UC UD - N C ∗∗ h = 0 SE N - / ∗∗ ∗ ∗ ∗ ∗ v = 0 C / - ∗∗ Table 9: Gauges-overview. One can choose two dierent gauge conditions. A slash denotes combina- tions that are not possible. For some important gauges we introduce names. Stars denote residual gauge freedom (see text for explanation). initial value for Φ. Other gauge-invariant quantities will be introduced later, in subsection 4.4.3.

4.2.5 Gauges

We can x the gauge by putting constrains on the metric perturbation elds or the uid perturbation elds. Henceforth, we will only consider scalar perturbations and neglect all vector and tensor per- turbations. This can be done since in rst order scalar, vector and tensor modes decouple. Then the following transformations are relevant:

φ˜ = φˆ − Hξ0 − ξ00, (4.45) ψ˜ = ψˆ + Hξ0, (4.46) w˜ =w ˆ + ξ0 − ξ0, (4.47) h˜ = hˆ − ξ, (4.48) v˜ =v ˆ + ξ0, (4.49) δ˜ = δˆ + 3Hξ0. (4.50)

Now we can choose two of these elds to be zero, which corresponds to choosing ξ0 and ξ. Note that not all combinations are possible. One can only choose h or v to be zero (which xes ξ), as well as one can only choose φ or ψ or δ to be zero (which xes ξ0). However, one can always choose w to be zero, because this only xes the dierence between ξ00 and ξ. In table 9 we show an overview over the possible gauges one can construct by combining constrains on these perturbation elds. A slash indicates a gauge that does not exist. For example, it is not possible to construct a gauge in which both ψ and φ are zero, because then eq. (4.45) and eq. (4.46) would give contradicting expressions for ξ0. The stars indicate the following residual gauge freedoms:

• * : There is a residual gauge freedom ξ → ξ + C(x), where C(x) is an arbitrary function.

39 ** : There is a residual gauge freedom 0 0 1 , where is an arbitrary function. • ξ → ξ + a D(x) D(x) Hence, gauges with one and/or two stars are not unique, and we have to x the residual gauge freedom (e.g. by setting C(x) = D(x) = 0). Furthermore, we have given some important gauges names in order to refer to them:

• UC - uniform curvature gauge. In this gauge we set ψ = w = 0. One can show that in a spatially at universe the 3-Ricci-scalar is given by (3)R = 4∆ψ (see [11]). Hence, in the uniform curvature gauge, as well as in any other gauge with ψ = 0 there is no intrinsic curvature. These gauges are hence a good choice for the comparison to Newtonian cosmology, where curvature does not exist.

• SE - spatially Euclidean gauge. In this gauge we set ψ = h = 0, so that again the intrinsic curvature vanishes. Furthermore, in this gauge the spatial part of the metric has no perturbation, i.e. looks Euclidean. All perturbations are in the (0, 0)-part and the (0, i)-part of the metric.

• UD - uniform density gauge. In this gauge there are no density perturbations, since we set δ = 0.

• N- conformal Newtonian gauge. In this gauge we set w = h = 0. It is equivalent to the zero shear gauge, σ = w = 0, where σ ≡ h0 + w generates the traceless part of the the extrinsic curvature tensor and hence can be interpreted as the shear in the normal worldlines [11]. Another common

name for it is longitudinal gauge. Note that the metric perturbations φN and ψN coincide with

the Bardeen potentials in this gauge, φN = Φ and ψN = Ψ. Furthermore, it can be shown that Φ corresponds to the gravitational potential introduced in Newtonian cosmology. For this, consider a static universe6. The line element in conformal Newtonian gauge has the form

2 µ ν 2 i j ds = gµν dx dx = −(1 + 2Φ)dτ + (1 − 2Ψ)δijdx dx . (4.51)

The proper time between two events along a worldline is given by

r Z p Z dxi dxj −ds2 = dτ 1 + 2Φ − (1 − 2Ψ)δ . (4.52) ij dτ dτ

Now we expand the integrand and keep only terms linear in Φ, Ψ and quadratic in v:

Z p Z v2 −ds2 = dt(1 + Φ − ). (4.53) 2

The proper time is extremized if and only if the Lagrangian v2 satises the Euler- L = 1 + Φ − 2 Lagrange equation ∂L d ∂L = , (4.54) ∂x dt ∂v which yields Newton's law for the motion of particles,

v˙ = −∇Φ. (4.55)

6A static universe has the nice simplications a = 1, dt = dτ, H = 0, u = v, x = r.

40 Figure 14: Illustration of slicing and threading, depending on the shift vector (here B) [8]. See text for explanation.

Hence, the Bardeen potential Φ can be identied with the gravitational potential.

• C- comoving gauge. In this gauge we set w = v = 0. To explain what this means more vividly, we need to introduce some new vocabulary. A gauge can be interpreted as a way of cutting the 4-dimensional spacetime into space and time. Hypersurfaces of constant τ give 3-dimensional

slices and worldlines of constant xi give 1-dimensional threads. The shift vector w tells how much the coordinates of a point are shifted from one slice to the next slice. The shift vector vanishes if and only if slicing and threading are orthogonal (see gure 14). The slicing is said to be comoving, if the time slices are always orthogonal to the uid 4-velocity, which can be achieved by choosing w = v. The threading is said to be comoving, if the threads are worldlines of comoving observers, i.e. v = 0. The comoving gauge is dened by requiring both comoving slicing and comoving threading [16, 8], which gives v = w = 0, or in the case of scalar perturbations, v = w = 0. Hence, in the comoving gauge slicing and threading are orthogonal and the threads are worldlines of comoving observers.

• S- synchronous gauge. In this gauge we set w = 0, so that comoving observers do not change their coordinates from one time-slice to the next one, and φ = 0, so that observers at dierent places have synchronous clocks (note that φ only aects the time-time component of the metric tensor).

41 4.3 Relativistic equations in the conformal Newtonian gauge

In the conformal Newtonian gauge, the metric tensor is

! 2 −1 − 2Φ 0 gµν = a (4.56) 0 (1 − 2Ψ)δij and the inverse metric is ! −1 + 2Φ 0 gµν = a−2 . (4.57) 0 (1 + 2Ψ)δij Using the metric tensor, we can construct the connection coecients,

1 Γµ ≡ gµλ(g + g − g ). (4.58) αβ 2 λβ,µ αλ,β αβ,λ A calculation gives:

0 0, 0 , 0 0 , Γ00 = H + Φ Γ0k = Φ,k Γij = (H − 2H(Φ + Ψ) + Ψ )δij i , i 0 , i i i . Γ00 = Φ,i Γ0j = (H − Ψ )δij Γkl = −(Ψ,lδk + Ψ,kδl ) + Ψ,iδkl

The Ricci tensor is dened as

α α α β α β (4.59) Rµν ≡ Γµν,α − Γαµ,ν + ΓαβΓµν − ΓβµΓαν .

The components are:

0 00 0 0 R00 = −3H + 3ψ + ∆Φ + 3H(Φ + Ψ ), 0 R0i = 2(Ψ + HΦ),i, 0 2 Rij = (H + 2H )δij 00 0 0 2 +[−Ψ + ∆Ψ − H(Φ + 5Ψ) − (2H + 4H )(Φ + Ψ)]δij

+(Ψ − Φ),ij.

Raising an index we nd

0 −2 0 −2 00 0 0 0 R0 = 3a H + a [−3ψ − ∆Φ + −3H(Φ + Ψ ) − 6H Φ], 0 −2 0 Ri = −2a (Ψ + HΦ),i, i −2 0 R0 = 2a (Ψ + HΦ),i, i −2 0 2 i Rj = a (H + 2H )δj −2 00 0 0 2 +a [−Ψ + ∆Ψ − H(Φ + 5Ψ) − (2H + 4H )(Φ + Ψ)]δij −2 +a (Ψ − Φ),ij.

42 The Ricci scalar is

0 i R ≡ R0 + Ri = 6a−2(H0 + H2) +a−2[−6Ψ00 + 2∆(2Ψ − Φ) − 6H(Φ0 + 3Ψ0) − 12(H0 + H2)Φ].

Finally, we can construct the Einstein tensor µ µ 1 µ, Gν ≡ Rν − 2 Rδν

0 −2 2 −2 0 2 G0 = −3a H + a [2∆Ψ + 6HΨ + 6H Φ], 0 0 −2 0 Gi = Ri = −2a (Ψ + HΦ),i, i i −2 0 G0 = R0 = 2a (Ψ + HΦ),i, 1 Gi = Ri − Rδi j j 2 j −2 0 2 i = a (−2H − H )δj −2 00 0 0 0 2 i +a [2Ψ + ∆(Φ − Ψ) + H(2Φ + 4Ψ ) + (4H + 2H )Φ]δj −2 +a (Ψ − Φ),ij.

The energy-momentum tensor for a perfect uid with zero pressure is

µ µ µ (4.60) Tν ≡ ρu uν =ρ ¯(1 + δ)u uν .

µ µ µ ν For the 4-velocity u we nd, using the normalization condition uµu = gµν u u = −1, ! 1 1 − Φ uµ = , (4.61) a vN,i ! −1 − Φ uµ = a . (4.62) vN,i

Note that vN = ∇vN , since we only consider scalar perturbations. Thus, the energy-momentum tensor up to linear order is: ! −ρ¯(1 + δ )ρv ¯ µ N N,i (4.63) Tν = . −ρv¯ N,i 0

Now we consider the Bianchi identity or relativistic energy-momentum conservation equation,

µ (4.64) ∇µTν = 0.

First consider the (ν = 0)-component of this equation. At background order, this gives the continuity equation that we know already from Newtonian physics,

ρ¯0 + 3Hρ¯ = 0, (4.65)

43 while the perturbed part gives the relativistic version of the perturbed continuity equation,

0 0 (4.66) δN + ∆vN − 3Ψ = 0.

Note the extra term −3Ψ0, which does not appear in Newtonian cosmology. Now consider the (ν = i)- component of eq. (4.64). There is no background order of this equation because ¯µ . However, the Ti = 0 perturbed part of the (ν = i)-component gives the relativistic version of the perturbed Euler equation,

0 (4.67) ∇vN + H∇vN = −∇Φ, which coincides exactly with the Newtonian equation. Now consider the Einstein eld equations,

µ µ (4.68) Gν = 8πGTν .

To the background order, both sides are diagonal, and we have two equations: The (0, 0)-component gives the Friedmann equation, 8πG H2 = ρa¯ 2, (4.69) 3 and the trace of the (i, j)-components gives the Raychaudhuri equation,

4πG H0 = − ρa¯ 2. (4.70) 3

At linear order however, there are 4 equations, the (0, 0)-component, the (0, i)-component, the trace of the (i, j)-components, and the traceless part of the (i, j)-components. These are in conformal Newtonian gauge:

2 0 2 3H Φ + 3HΨ − ∆Ψ = −4πGa ρδ¯ N , (4.71) 0 2 Ψ + HΦ = −4πGa ρv¯ N , (4.72) 3Ψ00 + 3H(Φ0 + 2Ψ0) + ∆(Ψ − Φ) + (2H0 + H2)Φ = 0, (4.73)

(Ψ − Φ),ij = 0. (4.74)

Note that the (0, 0)-component is the relativistic version of the Poisson equation, but the other equa- tions do not have any counterpart in Newtonian physics. Hence, in relativistic cosmological perturba- tion theory we have twice as much equations as in Newtonian cosmology.

4.4 Solutions

4.4.1 Solutions in the conformal Newtonian gauge

From eq. (4.74) we see that Φ and Ψ can only vary by some function of time, Φ − Ψ = f(τ). However, since the mean value of a Gaussian perturbation vanishes (see appendix B), we must have f(τ) = 0.

44 Thus, Φ = Ψ. (4.75)

Using this and the Raychaudhuri equation, eq. (4.73) becomes

Φ00 + 3HΦ0 = 0. (4.76)

The decaying solution is Φ ∼ τ −5 ∼ a−3/2 and the growing one is

Φ = Φ(x). (4.77)

Both the growing and the decaying mode coincide with the solution for the gravitational pertur- bation in the Newtonian case. From eq. (4.67) we nd the velocity potential,

2Φ v = − . (4.78) N 3H Again, this coincides with the velocity potential in the Newtonian case. Now consider the relativistic Poisson equation, eq. (4.71), in Fourier space. Using the Friedmann equation 2 3 2, 4πGa ρ¯ = 2 H Φ = Ψ and Ψ0 = 0, we obtain: 2k2 δ = −2Φ − Φ. (4.79) N 3H2 For scales that are small compared to the horizon, k , this corresponds to the Newtonian result. H  1 However, on scales larger than the horizon, k , the density contrast stays constant. H  1

4.4.2 Solutions in other gauges

We can construct solutions in other gauges using gauge transformations. For example, consider a gauge transformation from the conformal Newtonian gauge (wN = hN = 0) to the uniform curvature gauge (ψUC = wUC = 0). Equations (4.46) and (4.47) give

ψ Φ ξ0 = ξ0 = N = . (4.80) H H Then the density contrast in uniform curvature gauge is, according to eq. (4.50),

2k2 δ = δ − 3Hξ0 = δ − 3Φ = −5Φ − Φ. (4.81) UC N N 3H2

Note that we express the density contrast in terms of the Bardeen potential Φ, so that the Newtonian correspondence on small scales is satised. Hence, also in the uniform curvature gauge the density contrast stays constant outside the horizon, however the value of this constant diers by a factor of 5 . 2 The density contrast in other gauges can be constructed in the same way. An overview can be found in table 10, where we set all residual gauge modes to zero. Note that the density contrast in the comoving gauge and in the synchronous gauge have the same form as in Newtonian cosmology. 3/2 However, in other gauges the density contrast stays constant outside the horizon. Plots of ∆ = √k |δ| δ 2π2

45 gauge N UC SE S C UD 5 5 0 0* 5 k2 φ Φ 2 Φ 2 Φ − 9 H2 Φ 0 0 5 5 5 2 k2 ψ Φ 3 Φ 3 Φ 3 Φ + 9 H2 Φ 0 0 Φ 0 0 0 w − H 0 Φ 0 2 2 2 1 k2 h H2 − 3H2 Φ − 3H2 Φ − 3H2 Φ − 3 H4 Φ - 2 1 - 5 1 - 2 1 0* 0 2 k2 v 3 H Φ 3 H Φ 3 H Φ 9 H3 Φ 2 k2 2 k2 2 k2 2 k2 2 k2 0 δ −2Φ − 3 H2 Φ −5Φ − 3 H2 Φ −5Φ − 3 H2 Φ − 3 H2 Φ − 3 H2 Φ 2 k2 3ζ −5Φ − 3 H2 Φ Table 10: Overview - perturbations in conformal Newtonian (N), uniform curvature (UC), spatially Euclidean (UC), synchronous (S), comoving (C) and uniform density (UD) gauge. The quantity 3ζ = δ − 3ψ is gauge-invariant. (* This is a result of the calculations and not the gauge condition.) in dierent gauges can be found in gures 15 and 16. A discussion will follow in part 5.

4.4.3 Gauge-invariant solutions

In order to avoid the problem of gauge-dependence, one can consider only gauge-invariant perturba- tions. However, then the diculty arises in interpreting these quantities, which is only possible if one evaluates the gauge-invariant quantity in a specic gauge. As an example, consider the following three dierent gauge-invariant combinations containing the density contrast, all introduced by Bardeen [17]:

• The quantity δGI,UC = δ − 3ψ = 3ζ reduces to δ in gauges where ψ = 0. Hence, it measures the density contrast on hypersurfaces with zero curvature.

0 • The quantity δGI,N = δ−3H(w−h ) reduces to δ in a gauge where w = h = 0. Hence, it measures the density contrast on longitudinal hypersurfaces, i.e. hypersurfaces whose normal unit vectors have zero shear.

• The quantity δGI,C = δ −3H(v +w) reduces to δ in a gauge where v = w = 0. Hence, it measures the density contrast on comoving hypersurfaces, i.e. from the point of view of the matter [17].

Expressed in terms of the Bardeen potential, these quantities are

2 k2 δ = −5Φ − Φ, (4.82) GI,UC 3 H2 2 k2 δ = −2Φ − Φ, (4.83) GI,N 3 H2 2 k2 δ = − Φ. (4.84) GI,C 3 H2

All three quantities coincide on scales well within the horizon, k  H, but they dier signicantly outside the horizon. Which of these gauge-invariant quantities can be considered as the physically most appropriate is, as Bardeen formulated it, a matter of taste [17]. Thus, the introduction of gauge-invariant variables does not solve the gauge problem. Instead of choosing a gauge, one has to choose which gauge-invariant variable to take.

46 0.1

0.01 ) horizon entry -1 0.001

(a, k=0.01Mpc 1e-04 ∆

1e-05

δ δ C,S UC,SEδ N Φ 1e-06 1e-04 0.001 0.01 0.1 1 a

Figure 15: Evolution of the density contrast for k = 0.01 Mpc−1 in dierent gauges.

0.001

1e-04 )

-1 1e-05

(a, k=0.001Mpc 1e-06 ∆

1e-07 horizon entry

δ δ C,S UC,SEδ N Φ 1e-08 1e-04 0.001 0.01 0.1 1 a

Figure 16: Evolution of the density contrast for k = 0.001 Mpc−1 in dierent gauges.

47 4.5 Vector and tensor perturbations

So far we have ignored vector and tensor perturbations and considered only scalar perturbations. This is possible, because in rst order they decouple from each other. For the same reason we can choose the gauge condition for the scalar perturbations (this corresponds to choosing ξ0 and ξ) and the vector perturbations (this corresponds to choosing ξ⊥) independently. Note that pure tensor perturbations are per se gauge-invariant, so there is no gauge-choice for them.

As an example, we can choose a gauge with the scalar gauge condition w = h = 0 and the additional vector gauge condition h = 0. We call this gauge the Poisson gauge (P), in agreement with [9]. The gauge conditions can be reformulated in the form

i (4.85) ∇iwP = 0, ij (4.86) ∇ihP = 0.

Note that the scalar perturbations in the Poisson gauge are the same as the scalar perturbations in the conformal Newtonian gauge. However, in the Poisson gauge, we have some additional vector perturbation in the metric, ⊥ Hence, in this gauge we nd the same equations for wP = wP 6= 0. scalar perturbations as in the conformal Newtonian gauge, but we nd some additional equations for the vector perturbations, as shown in [9]. These are the vector parts of the Euler equation, the (0, i)-component and the (i, j)-component of the Einstein equations (we drop the index P now - all equations are given in Poisson gauge):

⊥ ⊥ 0 ⊥ ⊥ (4.87) (vi + wi ) + H(vi + wi ) = 0, 1 ∆w⊥ = 8πGa2ρ¯(v⊥ + w⊥), (4.88) 2 i i i  0 ⊥ ⊥ (4.89) w(i,j) + 2Hw(i,j) = 0.

From eq. (4.87) we nd a decaying solution for the vector perturbations,

⊥ ⊥ −1 (4.90) vi + wi ∼ a .

If there would have been some vector perturbations initially, they would have decayed, so our initial approach to neglect all vector perturbations from the beginning is justied. However, note that this argument only holds in rst order perturbation theory. At second order, there are couplings between scalar and vector perturbations [18].

The only pure tensor perturbation that occurs is the transverse-traceless part of hij. In Poisson gauge, we have T , since and . The only equation with a pure tensor part is the hij = hij h = 0 h = 0 transverse-traceless part of the (i, j)-component of the Einstein eld equations,

00 0 (4.91) hij + 2Hhij − ∆hij = 0.

48 Neglecting the Hubble friction, the solutions are gravitational waves,

i(ωτ+k·x) hij ∼ e , (4.92) with the dispersion relation ω ∂ω = = 1, (4.93) k ∂k which means that gravitational waves propagate with the speed of light. The Hubble drag term causes the amplitude of these gravitational waves to decrease with time due to the expansion of the universe.

5 Quantication of relativistic corrections

In this part we quantify the relativistic corrections in cosmological N-body simulations and the ob- served matter power spectrum. Furthermore, we discuss briey how the situation changes if one considers a ΛCDM model instead of an Einstein-de Sitter model.

5.1 Relativistic corrections to Newtonian simulations

In the conformal Newtonian gauge and in the uniform curvature gauge the relativistic corrections for the density perturbations on large scales and for early times become signicantly large. This can have an impact on cosmological N-body simulations, which use Newtonian instead of relativistic equations. These simulations typically start at z = 100 and the largest scales they simulate are about k−1 ' 1 Gpc. We dene the relative error of the Newtonian density contrast compared to the relativistic one,

δ − δ D(a, k) ≡ GR Newton , (5.1) δNewton where we x the initial value for all solutions according to eq. (3.94). Plots of D(a, k) can be found in gure 17. The relative error increases dramatically for large scales and early times: It is of the order ∼ 103% (∼4000% in the uniform curvature gauge and ∼2000% in the conformal Newtonian gauge) on the 1 Gpc-scale at z = 100. For scales below 10 Mpc (galaxy and cluster scales) and redshifts below z = 100, the error does not excess 1%. Hence, Newtonian simulations are more reliable on these scales. Some authors [19, 20, 21] argue that Newtonian simulations are exact at linear order, if one uses the following dictionary (one-to-one correspondence) between the simulated quantities and gauge- invariant relativistic quantities:

δφsim ↔ Φ, (5.2)

vsim ↔ vGI,N, (5.3)

δsim ↔ δGI,C, (5.4)

0 where vGI,N = v + h , so that it reduces the velocity potential on zero shear hypersurfaces. However, there is a problem arising with this dictionary: While there is one set of equations connecting the quantities on the left hand side (namely the Newtonian equations), there is not one set of equations

49 10000

1000

100

10

1 [%] UC

D 0.1

0.01

0.001

-1 1e-04 k=1Mpc k=0.1Mpc-1 k=0.01Mpc-1 k=0.001Mpc-1 1e-05 0.01 0.1 1 a

10000

1000

100

10

1 [%] N D 0.1

0.01

0.001

-1 1e-04 k=1Mpc k=0.1Mpc-1 k=0.01Mpc-1 k=0.001Mpc-1 1e-05 0.01 0.1 1 a

Figure 17: The relative error D(a, k) in the uniform curvature gauge (top) and the conformal Newtonian gauge (bottom). 50 1e+06

100000 horizon ] 3 10000 [Mpc δ P

1000

C,S UC,SE N 100 1e-04 0.001 0.01 0.1 1 k [Mpc-1]

Figure 18: The matter power spectrum today in dierent gauges. connecting the quantities on the right hand side. Indeed, this dictionary identies the simulated grav- itational and velocity potentials with the relativistic counterparts on zero shear hypersurfaces, but the simulated density contrast with the relativistic one on comoving hypersurfaces. This means that New- tonian simulations are practically a mixture of relativistic simulations on two dierent hypersurfaces.

Furthermore, note that we cannot observe δGI,C since we are not observers on a comoving hypersurface. A consistent dictionary would be the following,

δφsim ↔ Φ, (5.5)

vsim ↔ vGI,N, (5.6)

δsim ↔ δGI,N + 2Φ, (5.7) where we identify all simulated quantities with gauge-invariant quantities that have a physical inter- pretation on zero shear hypersurfaces.

5.2 Relativistic corrections to the observed power spectrum

Another interesting result is the rise of the matter power spectrum at very large scales, see gure 18. At scales near the horizon today the relativistic power spectrum in the uniform curvature gauge is about two orders of magnitude higher than in the the comoving/synchronous gauge. This rise of the

51 −1 z kmin [Mpc ] 1 1.25 · 10−3 0.6 1.75 · 10−3 0.3 2.98 · 10−3 Table 11: The minimal observable wavenumber for various redshifts. power spectrum at large scales can be misinterpreted as primordial non-Gaussianity in the gravitational potential perturbation, as pointed out by Yoo in [22] and by Zhang in [23]. As a next step we want to nd out if this is an observable eect in the measured power spectrum at a given redshift. For this, we introduce the distance-redshift relation in an Einstein-de Sitter universe [4],

2  1  d(z) = 1 − √ . (5.8) H0 1 + z

The minimal wavenumber one can observe at a given redshift is given by the inverse of d(z),

H0 π kmin = , (5.9) 2 1 − √ 1 1+z where the extra factor π comes from the fact that we want to measure at least half a wavelength (note that the relation between wavenumber and wavelength is 2π ). Table 11 shows for the k = λ kmin redshifts z = 0.3, z = 0.6 and z = 1. In gure 19 we show the matter power spectra for these redshifts, where the vertical bar denotes kmin. The dierence due to the gauge choice can become signicantly large: At a redshift of z = 1, the power spectrum at the minimal observable wavenumber is about twice as high in the uniform curvature gauge as it is in the comoving gauge.

2 For completeness we also show the velocity power spectrum Pv(k) = |vk| today in dierent gauges in gure 20. In the conformal Newtonian and the spatially Euclidean gauge it corresponds to the Newtonian power spectrum, see gure 12. However, in the uniform curvature gauge the spectrum is higher by a factor of 25 . In the uniform density gauge the spectrum has a totally dierent shape. 4 Thus, the property that all gauges coincide on sub-horizon scales, which we found in the case of density perturbations, is not true for velocity perturbations. Note that in the comoving gauge and in the synchronous gauge the spectrum is zero because the peculiar velocity vanishes in these gauges, see table 10.

5.3 Relativistic corrections in ΛCDM models

As we approach the present time, the cosmological constant gets more and more dominant. Today we have ΩΛ ' 0.7, so that we should not neglect the eect of Λ. We can include dark energy into our equations by setting and , where Λ const and Λ . ρ = ρm + ρΛ p = pm + pΛ ρΛ = 8πG = . pΛ = −ρΛ = − 8πG Then the modied energy-momentum tensor is

µ µ µ (5.10) Tν = (ρ + p)u uν + pδν .

52 10000 ] 3 (z=1) [Mpc δ P

1000

C,S UC,SE N 0.001 0.01 k [Mpc-1]

10000 ] 3 (z=0.6) [Mpc δ P

C,S UC,SE N 1000 0.001 0.01 k [Mpc-1]

10000 ] 3 (z=0.3) [Mpc δ P

C,S UC,SE N 1000 0.001 0.01 k [Mpc-1]

Figure 19: The matter power spectrum between k = 0.001 Mpc−1 and k = 0.01 Mpc−1 in dierent gauges for the redshifts z = 1, z = 0.6 and z = 0.3. The vertical line denotes kmin, the minimal observable wavenumber at the corresponding redshift.

53 10000

100

1

0.01

1e-04 ] 3 1e-06 [Mpc v P 1e-08

1e-10

1e-12

1e-14 N,SE UC UD 1e-16 1e-04 0.001 0.01 0.1 1 k [Mpc-1]

Figure 20: The velocity power spectrum today in dierent gauges.

Note that and Λ , so that we have ρ + p = ρm + ρΛ + pm + pΛ = ρm p = pm + pΛ = − 8πG

µ µ µ (5.11) Tν = ρmu uν − ρΛδν .

Hence, the only change is in the background,

! −ρ¯ − ρ 0 ¯µ m Λ (5.12) Tν = , 0 −ρΛδij while the perturbations remain unchanged,

! −ρ¯ δ ρ¯ v µ m N m N,i (5.13) δTν = . −ρ¯mvN,i 0

Now we calculate the corrections that appear in the background equations. First, consider the Bianchi identity µ . For we found the mass continuity equation, ∇µTν = 0 ν = 0

0 (5.14) ρ¯m + 3Hρ¯m = 0, which remains unchanged, since Λ is constant in time. The (ν = i)-component of the Bianchi identity still has no background part, which comes from the fact that Λ is constant in space. Now consider the Einstein eld equations, µ µ. In the background, the -component gives the Friedmann Gν = 8πGTν (0, 0)

54 equation, which becomes now 8πG H2 = a2(¯ρ + ρ ). (5.15) 3 m Λ The Raychaudhuri equation (trace of the (i, j)-components) becomes

0 2 2 2H + H = 8πGa ρΛ, (5.16) or, using the Friedmann equation,

4πG H0 = − a2(¯ρ − 2ρ ). (5.17) 3 m Λ In terms of proper time the background equations are

ρ¯˙m + 3Hρ¯m = 0, (5.18) 8πG H2 = (¯ρ + ρ ), (5.19) 3 m Λ a¨ 4πG = − (¯ρ − 2ρ ). (5.20) a 3 m Λ One can nd solutions to these background equations in terms of hyperbolic functions [24],

 Ω 1/3  t  a(t) = m sinh2/3 , (5.21) 1 − Ωm tΛ 2  t  H(t) = coth , (5.22) 3tΛ tΛ   −2 t ρ¯m(t) = ρΛ sinh , (5.23) tΛ

2 where tΛ = √ . Now consider the perturbed equations in conformal Newtonian gauge: 24πGρΛ

δ0 − k2v − 3Φ0 = 0, (5.24) v0 + Hv = −Φ, (5.25) 2 0 2 2 3H Φ + 3HΦ + k Φ = −4πGa ρ¯mδ, (5.26) 0 2 Φ + HΦ = −4πGa ρ¯mv, (5.27) 00 0 2 Φ + 3HΦ = −8πGa ρΛΦ. (5.28)

These are the perturbed parts of the 0-component and the i-component of the energy-momentum conservation equation, the (0, 0)-component, (0, i)-component and (i, i)-component of the Einstein eld equations, respectively, where we already used Φ = Ψ, which follows from the traceless part of the (i, j)-components, which remains unchanged. Note that the only change occurs in the (i, i)- component, eq. (5.28), because the term 2H0 +H2 is not zero any more as we leave matter domination. An analytic solution of these equations is very dicult due to the non-triviality of the background solutions. However, one can do a numerical integration of the above equations to nd the evolution of

55 Figure 21: Relative error D(z, k) (here denoted as ∆) between the relativistic density contrast in conformal Newtonian gauge and the Newtonian one. From top to bottom, the curves are for k = 0.001 Mpc−1, k = 0.01 Mpc−1, k = 0.1 Mpc−1 and k = 1 Mpc−1, respectively [25]. the density contrast. This was done for example by Dent and Dutta in [25], where they also solved the corresponding Newtonian equations numerically and plotted the relative error between the Newtonian solution and the relativistic solution. The result is shown in gure 21: The relative errors become even larger (∼ 104% today on 1 Gpc-scales) in ΛCDM models than in the Einstein-de Sitter model (where we found at most ∼ 103% on these scales). Note that the relative error remains constant in time due to the fact that the density perturbations freeze out as dark energy takes over.

56 6 Summary and outlook

In this thesis we studied the validity of Newtonian equations, as used in cosmological N-body simula- tions at large cosmological scales. These simulations are used to test our understanding of the universe, in particular the formation of galaxies and clusters of galaxies. We linearized these equations in order to compare them to linear relativistic cosmological perturbation theory. The linear equations need to be corrected on small scales due to nonlinear eects and on large scales due to relativistic eects. The important results are the following:

• For typical density uctuations (≤ 1σ) linear theory is valid until the density contrast reaches

unity. This happened at a scale factor of aNL2 = 0.12 (corresponding to a redshift of zNL2 ' 7) on

galaxy scales and aNL2 = 0.75 (zNL2 ' 0.3) on cluster scales. Perturbations on supercluster scales still behave linear today (all second order quantities are smaller than all rst order quantities).

• The behaviour of density perturbations on scales larger than the Hubble horizon is strongly gauge- dependent. It turns out that in the synchronous and the comoving gauge the density contrast behaves like in Newtonian theory. However, in the conformal Newtonian gauge, the spatially Euclidean gauge and the uniform curvature gauge the density contrast is constant outside the Hubble horizon.

• The relativistic corrections to the simulated density perturbations in Newtonian N-body sim- ulations can become signicantly large. Assuming that the accurate description of the density contrast on superhorizon scales is given in the uniform curvature gauge or the conformal New- tonian gauge, the error of the density contrast in Newtonian simulations reaches up to ∼ 103% on the 1 Gpc-scale at a redshift of z = 100. This error increases to ∼ 104% if we consider a ΛCDM model instead of an Einstein-de Sitter model. For scales below 10 Mpc and redshifts below z = 100, the error does not excess 1%, whence Newtonian simulations on these scales are more reliable.

• The observed matter power spectrum diers on large scales in dierent gauges. For example, at a redshift of z = 1, the power spectrum at the minimal observable wavenumber is about twice as large in the uniform curvature gauge as it is in the comoving gauge. It remains an open question which power spectrum we expect to measure from a theoretical point of view. Maybe the gauge-dependence of the power spectrum is a cue that this is not at all the quantity we should consider.

• We should not ask which gauge is the correct one. Indeed, perturbations in all gauges are correct solutions of Einstein's equations - this is the basic principle of . Instead, we should ask the following questions: Is there a chosen gauge which gives predictions that are closest to our measurements? If yes, which gauge and why? Another possibility is that there is no such chosen gauge and indeed all quantities that we measure are truly gauge-invariant quantities, which we do not yet understand theoretically. We have to answer these questions rst before we really know how reliable Newtonian N-body simulations at large cosmological scales are.

57 Appendix A Helmholtz decomposition theorem

Bounded domain

Let v(r) be a smooth vector function on a bounded domain D ⊂ R3, v : D → R3. Then there exists a scalar potential v and a vector potential Ω such that:

v = vk + v⊥, (A.1) = ∇v + ∇ × Ω, (A.2)

with vk =∇v and v⊥ = ∇ × Ω, ∀ r ∈ D. It follows that ∇ · v⊥ = 0 and ∇ × vk = 0. The proof is done by construction. One can solve

∇ · vk = ∇ · v, (A.3)

∇ × v⊥ = ∇ × v, (A.4) using the relation 1 ∆ = −4πδ(r − r0). (A.5) |r − r0| The solutions are given by

Z 0 0 Z 0 1 3 0 ∇ · v(r ) 1 0 v(r ) (A.6) v = − d r 0 + dS · 0 , 4π D |r − r | 4π ∂D |r − r |

Z 0 0 Z 0 1 3 0 ∇ × v(r ) 1 0 v(r ) (A.7) Ω = d r 0 − dS × 0 . 4π D |r − r | 4π ∂D |r − r |

Unbounded domain

If the domain D is unbounded, i.e. D = R3, we have to make the restriction

|v(r)| → 0 as |r| → ∞, (A.8) |r| so that the second terms in v and Ω vanish. Then the solutions are given by

Z 0 0 1 3 0 ∇ · v(r ) (A.9) v = − d r 0 , 4π D |r − r |

Z 0 0 1 3 0 ∇ × v(r ) (A.10) Ω = d r 0 . 4π D |r − r |

Hypertorus

Now let the domain D be a Hypertorus, D = T3, as used in numerical simulations, where one considers a box with periodic boundary conditions. Since we have v(ra) = v(rb) and dS(ra) = −dS(rb) for two

58 opposite sites a and b of the hypertorus, the boundary terms in v and Ω vanish as in the case of the unbounded domain.

Uniqueness

The Helmholtz decomposition is not unique. If the vector contains a constant part, which is both divergence- and curl-free, it can be included in the longitudinal part or in the transverse part. Note that in cosmology a constant vector stands in contradiction to the isotropy of the universe. In order to have a unique decomposition, one has to assume natural boundary conditions, i.e. v = 0 at the boundary of the domain. In particular, in the case of an unbounded domain, the boundary condition is v(|r| → ∞) → 0. (A.11)

For consistency, we show that the peculiar velocity we found satises this condition in all gauges con- −1/2 sidered. We found |vk| ∼ kΦ ∼ k T (k) in the conformal Newtonian gauge, the uniform curvature 3 1/2 gauge and the spatially Euclidean gauge and |vk| ∼ k Φ ∼ k T (k) in the uniform density gauge (see n table 10). Now we will consider the general case |vk| ∼ k T (k) and show that its Fourier transform vanishes for |x| → ∞ for all n ∈ R. We have Z 3 ik·x |v(x)| ∼ d k|vk|e Z ∼ d3kknT (k)eik·x Z ≤ d3kkneik·x

Z kUV Z π ∼ dk dθ sin θkn+2eikx cos θ kIR 0 Z kUV sin(kx) ∼ dkkn+2 , kIR kx where we introduced the ultraviolet cuto wavenumber kUV and the infrared cuto wavenumber kIR (this is a reasonable assumption for a Harrison-Zel'dovich spectrum, see appendix B). Now we make the substitution p ≡ kx, so that

Z xkUV 1 n+2 sin(p) |v(x)| ∼ n+3 dpp x xkIR p Z xkUV 1 n+1 ≤ n+3 dpp x xkIR 1 1 = xn+2(kn+2 − kn+2) xn+3 n + 2 UV IR 1 ∼ . x

59 for n 6= −2. For n = −2 we have

1 Z xkUV |v(x)| ≤ dpp−1 x xkIR 1 k  = ln UV x kIR 1 ∼ . x

Thus, the boundary condition v(|x| → ∞) → 0 is satised in all gauges we considered in this thesis.

Appendix B Statistical properties of Gaussian perturbations

A Gaussian perturbation Z g(x) = d3kg(k)eik·x (B.1) is dened as a perturbation whose Fourier components have a Gaussian probability distribution,

2 Prob 1 |g(k)| (B.2) [g(k)] = 2 exp(− 2 ), 2πsk 2sk where 2 is the variance. From this we see directly that the mean value of each Fourier component is sk zero, hg(k)i = 0. (B.3)

Furthermore, if we assume homogeneity and isotropy of the perturbations, it follows that dierent Fourier components satisfy the following normalization condition:

hg(k)g(k0)i = δ(k + k0)P (k), (B.4) where P (k) ≡ |g(k)|2 (B.5) is the power spectrum7. Now we go back to coordinate space and nd the following relation:

Z hg(x)i = h d3kg(k)eik·xi Z = d3khg(k)ieik·x = 0.

7This commonly used denition of the power spectrum is formally wrong. Indeed, one should divide by a unit volume, so that the power spectrum has the correct dimension, namely Mpc3. However, this is only a formal issue, so this unit volume is usually dropped.

60 Hence, the mean value of a Gaussian perturbation in coordinate space vanishes. This comes from the fact that positive and negative deviations from the background value are equally probable. In other words, the mean value of a perturbed quantity is its background quantity. Next, consider

3 3 0 Z d k Z d k 0 hg(x)g(y)i = h eik·x eik ·yg(k)g(k0)i (2π)3 (2π)3 Z d3k = eik·(x−y)P (k) (2π)3 Z ∞ k2dk sin(kr) = 2 P (k) 0 2π kr Z ∞ dk sin(kr) k3 = 2 P (k), 0 k kr 2π | {z } ≡P(k) where r ≡ |x − y| is the distance and

k3 k3 P(k) ≡ P (k) = |g(k)|2 (B.6) 2π2 2π2 is the dimensionless (band) power spectrum. For r = 0 we have

Z ∞ hg(x)2i = d ln kP(k). (B.7) −∞

This means that the power spectrum P(k) of g gives the contribution of a logarithmic scale interval to the variance of g(x).A power-law spectrum is a power spectrum of the form

k P(k) = A2( )n−1, (B.8) kpivot where kpivot is a chosen reference scale, the pivot scale, A is the amplitude at this scale, and n is the spectral index. For n = 1, the spectrum is scale-invariant,

P(k) = A2. (B.9)

This spectrum is also called the Harrison-Zel'dovich spectrum [8, 13]. Note that hg(x)2i is logarith- mically divergent for a Harrison-Zel'dovich spectrum. Therefore one usually denes a cuto-scale at small scales (ultraviolet-cuto) and at large scales (infrared-cuto) in order to make the integral nite.

61 Appendix C Perturbations in the inaton eld

For completeness, we consider here perturbations in the inaton eld, that lead to the important curvature perturbations after ination. We start with the Lagrangian for the inaton eld:

1 L = − ϕ,µϕ − V (ϕ). (C.1) 2 ,µ Applying the Euler-Lagrange equations, we nd the eld equation for the inaton,

2 (C.2) ϕ¨ + 3Hϕ˙ + ∇rϕ + V,ϕ = 0.

Now we split the inaton eld into a background part and a perturbed part,

ϕ(r, t) =ϕ ¯(t) + δϕ(r, t). (C.3)

Then, in zeroth order, the eld equation becomes

ϕ¯¨ + 3Hϕ¯˙ + V,ϕ = 0, (C.4) which we already found in section 2.3. In linear order the eld equation reads

¨ ˙ 2 (C.5) δϕ + 3Hδϕ − ∇rδϕ + δϕV,ϕϕ = 0.

In slow-roll approximation, i.e. V 2, we can neglect the last term on the left hand side. V,ϕϕ  2 ' H MP l The resulting equation becomes in Fourier space:

k2 δϕ¨ + 3Hδϕ˙ + δϕ = 0, (C.6) k k a2 k where k is the comoving wavenumber of the considered scale. Now we can distinguish two cases:

For early times, when k2 , that means before horizon exit, the inaton perturbation is • 3H  a2 oscillating with a decaying amplitude, 1 . |δϕk| ∼ a For late times, when k2 , that means after horizon exit, the solution is , or in • 3H  a2 δϕk = const. other words, δϕk freezes out.

Note that, because H ≈ const. during ination, all scales have similar behaviour. This leads to a nearly scale-invariant power spectrum of inaton-, and hence, curvature perturbations8. Indeed, one can even show that the spectral index is related to the slow roll parameters, as dened in section 2.3,

n − 1 = −6 + 2η, (C.7)

8Bardeen showed in [11] that curvature perturbations are related to inaton perturbations by δϕ Hence, ζ = −H ϕ˙ . their spectra dier only by a factor H2 , which is constant in the slow-roll approximation. ϕ˙ 2

62 whence n ' 1 in slow-roll approximation [8].

Appendix D Meszaros equation

The Meszaros equation gives the growth of cold dark matter density perturbations in the intermediate regime between radiation domination and matter domination. Assume that the universe is lled with cold dark matter and radiation to a comparable amount. Then the Friedmann and Raychaudhuri equations are:

8πG H2 = (¯ρ +ρ ¯ ), (D.1) 3 m γ a¨ 4πG = − (¯ρ + 2¯ρ ), (D.2) a 3 m γ where ρ¯m is the (dark) matter density and ρ¯γ is the radiation density. It is helpful to introduce a new time variable a ρ¯ y ≡ = m , (D.3) aeq ρ¯γ so that

y  1 ⇔ radiation domination, y  1 ⇔ matter domination.

For cold dark matter perturbations the Jeans equation gives

¨ ˙ δ + 2Hδ − 4πGρ¯mδ = 0. (D.4)

Note that there is no pressure source term because we consider only cold dark matter, which does not interact with radiation. Now we want to write this equation in terms of the new time variable y. To do this, we use 3 y 4πGρ¯ = H2 , (D.5) m 2 1 + y and the transformation of the time derivatives,

d dy d d = = Hy . (D.6) dt dt dy dy

After some calculation one nds the Jeans equation in terms of the new time variable,

d2δ 2 + 3y dδ 3 + − δ = 0. (D.7) dy2 2y(y + 1) dy 2y(y + 1)

63 This equation was rst derived by Meszaros [26] and hence carries his name, the Meszaros equation9. The general solution is given by

δ(y) = C1δ1(y) + C2δ2(y), (D.8) where C1 and C2 are constants and 3 δ1(y) = 1 + y, (D.9) 2 √  3  1 + y + 1 δ (y) = 3p1 + y − 1 + y ln √ . (D.10) 2 2 1 + y − 1

We can reconstruct the solutions during radiation domination and matter domination by considering the limits y  1 and y  1, respectively:

• During radiation domination (y  1) the leading order in δ2 is δ2 ∼ ln y ∼ ln a, and the leading

order in δ1 is a constant. Hence, δ2 is the growing mode and δ1 is the decaying mode.

• During matter domination (y  1) the leading order in δ1 is δ1 ∼ y ∼ a, and the leading order −3/2 −3/2 in δ2 is δ2 ∼ y ∼ a . Thus, now δ1 is the growing mode and δ2 is the decaying mode.

Appendix E Identity of synchronous and comoving gauge

We claim that, neglecting residual gauge modes, the synchronous gauge, where φS = wS = 0, and the comoving gauge, where vC = wC = 0, are identical gauges during matter domination in that all perturbation variables are the same in both gauges. To see that, rst consider a gauge transformation from a generic gauge G to the conformal Newtonian gauge N, where wN = hN = 0, which is given by

0 ξ = hG − wG, (E.1)

ξ = hG. (E.2)

This transformation is unique. Now consider a gauge transformation from the conformal Newtonian gauge to another gauge where two of the triple (φ, w, v) are zero. The transformation rules are:

φ = Φ − Hξ0 − ξ00, (E.3) w = ξ0 − ξ0, (E.4) 0 v = vN + ξ . (E.5)

9It is also possible to construct the Meszaros equation for the intermediate time interval between matter domination and dark energy domination. However, this equation is dicult to solve analytically. Böhmer et al. constructed a w-Meszaros equation for the time between matter domination and domination of a component with arbitrary w, and they were only able to nd two additional analytic solutions in the parameter space , one for 1 and w ∈ [−1, 1] w = − 3 one for 1 [27]. w = 6

64 0 Note that the transformation law for φ implies ξ = −vN ⇔ φ = 0, which follows from the Euler equation during matter domination. Now there are three possibilities:

Setting (synchronous gauge) implies 0 0 C , whence C . We can set • φ = w = 0 ξ = ξ = −vN + a v = a the constant to zero, so that v = 0.

0 0 • Setting v = w = 0 (comoving gauge) implies ξ = ξ = −vN , whence also φ = 0.

Setting implies 0 C and 0 , whence C . Again, we can set the • φ = v = 0 ξ = −vN + a ξ = −vN w = a constant to zero, so that w = 0.

Since all three gauges, in which two of the triple (φ, w, v) are zero, are achieved by choosing ξ0 = ξ0 =

−vN , they are identical. In particular, the comoving and the synchronous gauge are identical during matter domination. Note that this only holds if we set all residual gauge modes to zero.

65 References

[1] C.L. Bennett et al., Four year COBE DMR cosmic microwave background observations: Maps and basic results, Astrophys.J. 464:L1, 1996.

[2] J.R. Gott et al., A map of the universe, Astrophys.J. 624:463, 2003.

[3] V. Springel et al., Simulations of the formation, evolution, and clustering of galaxies and quasars, Nature 435:629, 2005.

[4] V.F. Mukhanov, Physical Foundations of Cosmology, Cambridge University Press 2005.

[5] A.H. Jae, Cosmology (lecture notes), Imperial College London 2011.

[6] K. Freese, Observational Evidence for Dark Matter, Talk given at conference: Dark Matter and Dark Energy, Lyon 2008, arXiv:0812.4005.

[7] S. Perlmutter, Status of the cosmological parameters, Int.J.Mod.Phys. A15S1:715, 2000.

[8] A.H. Liddle and D.H. Lyth, Cosmological Ination and Large-Scale Structure, Cambridge Uni- versity Press 2000.

[9] E. Bertschinger, Cosmological Dynamics, in: Les Houches Session LX: Cosmology and Large Scale Structure, North-Holland 1993, arXiv:9503125.

[10] D.N. Spergel et al., First year Wilkinson Microwave Anisotropy Probe (WMAP) observations: Determination of cosmological parameters, Astrophys.J.Suppl. 148:175, 2003.

[11] J. M. Bardeen, Cosmological perturbations: From quantum uctuations to large scale structure, Lectures given at 2nd Guo Shou-jing Summer School on Particle Physics and Cosmology, Nanjing, China, Jul 1988.

[12] M. Tegmark, A.J.S. Hamilton and Y.-Z. Xu, The power spectrum of galaxies in the 2dF 100k redshift survey, Mon.Not.Roy.Astron.Soc. 335:887, 2002.

[13] P.J.E. Peebles, The Large-Scale Structure of the Universe, Princeton University Press 1980.

[14] J.M. Bardeen, J.R. Bond, N. Kaiser and A.S. Szalay, The statistics of peaks of gaussian random elds, Astrophys.J. 304:15, 1986.

[15] V.F. Mukhanov, H.A. Feldman and R.H. Brandenberger, Theory of cosmological perturbations, Phys.Rept. 215:203, 1992.

[16] H. Kurki-Suonio, Cosmological Perturbation Theory (lecture notes), University of Helsinki 2010.

[17] J.M. Bardeen, Gauge invariant cosmological perturbations, Phys.Rev. D22:1882, 1980.

[18] J.-C. Hwang and H. Noh, Second-order perturbations of cosmological uids, Phys.Rev. D76:103527, 2007.

66 [19] N.E. Chisari and M. Zaldarriaga, Connection between Newtonian simulations and general rela- tivity, Phys.Rev. D83:123505.

[20] S.R. Green and R.M. Wald Newtonian and relativistic cosmologies, arXiv:1111.2997, 2011.

[21] J.-C. Hwang H. and Noh, Newtonian versus relativistic nonlinear cosmology, Gen.Rel.Grav 38:703, 2006.

[22] J. Yoo, General relativistic description of the observed galaxy power spectrum: Do we understand what we measure?, Phys.Rev. D82:083508, 2010.

[23] P. Zhang, The exact analytical solution of the linear structure growth rate in LCDM cosmology and its cosmological applications, Phys.Rev. D82:063510, 2011.

[24] O. Gron, A new standard model of the universe, Eur.J.Phys. 23:135, 2002.

[25] J.B. Dent and S. Dutta, On the dangers of using the growth equation on large scales, Phys.Rev. D79:063516.

[26] P. Meszaros, Primeval black holes and galaxy formation, Astron.Astrophys. 38:5, 1975.

[27] C.G. Boehmer and G. Caldera-Cabral, Evolution of sub-horizon cold dark matter perturbations, arXiv:1008.2852.

67 Acknowledgments

Firstly I want to thank my supervisor Prof. Dr. Dominik Schwarz for his great guidance, motiva- tion and ideas during my research. Furthermore, I want to thank Chris Byrnes for his help on the comprehension of the Mukhanov-paper, Matthias Rubart for his help on general cosmological ques- tions, and Viktor Dick for his help on some tricky calculations concerning the Christoel symbols and more. I also want to thank the whole theoretical physics work group, especially the cosmology work group, for creating an enjoyable working atmosphere. Last but not least, I want to thank my friends, especially Thomas Weiÿ and Thomas Hederer, and my family for their support during the last year.

68 Erklärung

Hiermit erkläre ich, dass ich die vorliegende Arbeit selbstständig verfasst und keine anderen als die angegebenen Hilfsmittel verwendet habe. Bielefeld, im Januar 2012.

(Samuel Flender)

69