<<

Universality of the spherical collapse with respect to the matter type : the case of a barotropic fluid with linear equation of state

Fran¸coisStaelens1, J´er´emy Rekier2, Andr´eF¨uzfa1∗ 1Namur Institute for Complex Systems (naXys), University of Namur, Belgium 2Royal Observatory of Belgium, Uccle, Belgium (Dated: April 15, 2021) We study the spherical collapse of an over-density of a barotropic fluid with linear equation of state in a cosmological background. Fully relativistic simulations are performed by using the Baumgarte- Shibata-Shapiro-Nakamura formalism jointly with the Valencia formulation of the hydrodynamics. This permits us to test the universality of the critical collapse with respect to the matter type by considering the constant equation of state parameter ω as a control parameter. We exhibit, for a fixed radial profile of the energy-density contrast, the existence of a critical value ω∗ for the equation of state parameter under which the fluctuation collapses to a and above which it is diluting. It is shown numerically that the mass of the formed black hole, for subcritical solutions, obeys a scaling law M ∝ |ω − ω∗|γ with a critical exponent γ independent on the matter type, revealing the universality. This universal scaling law is shown to be verified in the empty Minkoswki and de Sitter space-times. For the full matter Einstein-de Sitter universe, the universality is not observed if conformally flat sub-horizon initial conditions are used. But when super-horizon initial conditions computed from the long-wavelength approximation are used, the universality appears to be true.

I. INTRODUCTION tional waves, the signal GW150914 from a binary black holes merger in 2016 by [16] is an important evidence in favour of GR. This was made possible thanks to numer- The development of this 3+1 formalism of General Rel- th ical relativity which permitted to verify post-Newtonian ativity and the associated algorithmics during the XX analytical developments and to go beyond it by simulat- century combined with the ”computer revolution” of the ing the black holes merger. This has even enforced last decades permits the study of gravitation from a new Numerical Relativity as an active, powerful and essential point of view. The important works of Darmois ([1]), branch of physics. Lichnerowicz ([2], [3], [4]), Choquet-Bruhat ([5]), Dirac A second major result obtained thanks numerical rel- ([6], [7]) and Arnowitt et al. ([8]), managed to write ativity is the discovery of critical phenomena in gravita- Einstein equations as a constrained Cauchy problem, a tion. In 1992, Choptuik studied (see [17]) the spherical suitable form for numerical integration, in the view of gravitational collapse of a massless scalar field thanks developing a quantum theory of gravitation. These laid to numerical relativity. He found that, in a Minkowski the foundations of the Hamiltonian formulation of GR background, the mass of a formed black hole M follows and introduced the well known Arnowitt-Deser-Misner a scaling power-law (ADM) formalism. With the emergence of more and more powerful computers, numerical integration of such γ M ∝ (k − k∗) , (1) new formalisms became possible and the capability to simulate GR whatever the ingredients considered was where k is a one dimensional quantity parametrising the a dream that scientists could then try to render real- initial data, k∗ is the threshold for black holes formation istic. This opened the era of Numerical Relativity, a new (which means that a black hole is formed when k > k∗

arXiv:1912.00677v2 [gr-qc] 14 Apr 2021 field of research whose aim is to build and use numerical and not if k < k∗) and γ is the critical exponent which methods to solve Einstein equations of GR on a com- does not depend on k. Critical phenomena following such puter. Remarkable works to mention are [9], [10], [11], a scaling law are said ”universal”. Moreover, the critical [12], which developed the famous Baumgarte-Shibata- solution admits a continuous self-similarity (CSS). This Shapiro-Nakamura (BSSN) formalism in the 1990’s, the critical phenomenon is similar to critical phase transi- one used in this article. tions found in statistical mechanic by identifying M to Numerical Relativity obtained several successes and is an order parameter controlled by the function |k − k∗| now widely used in modern physics. It has been used, on the total phase space. On this basis, numerous exam- among others, to simulate spherical black holes forma- ples of universality in critical phenomena in gravitational tion ([13]), stable solutions of neutron stars ([14]) or bi- collapse were discovered, some with a critical solution ad- nary black holes ([15]),. . . The first detection of gravita- mitting a CSS and others with a discrete self-similarity (DSS). The interested reader can find more informa- tions in the review by Gundlach and Mart´ın-Garc´ıa[18]. Among others, it has been shown, still in a Minkowski ∗Electronic address: [email protected] background ([19], [20]), that the universality was true in 2 the case of the spherical collapse of a perfect fluid with ical coordinates are employed, it induces terms of the barotropic equation of state p = ωe, where p is the pres- form 1/rm which become problematic near the the ori- sure, e is the energy density of the fluid and ω is a con- gin of coordinates (i. e. when r → 0). To overcome this stant in the interval [0, 1]. The associed critical solution difficulty, the authors of [30] developped a partially im- is sometimes called the ”Evans-Coleman” CSS solution, plicit Runge-Kutta (PIRK) method for hyperbolic wave- according to the authors of [19]. It was unclear if such like equations which solves the problems of instabilities CSS solutions exist for ω > 0.89 until [21] showed it was without other regularization. This scheme has already the case for all ω between 0 and 1. This discovery was been applied with success in the case of asymptotic flat- of great importance because it means that, by fine tun- ness in [31]. For the case of an expanding background ing the initial conditions, it is possible to obtain a black universe, it has also been done but only in the case of hole with a mass as small as wished from a radiation dust in [24] and in the case of a scalar field in [25]. We fluid. The possible existence of tiny black holes would apply here to the case of pressured matter. thus have an impact on the aboundance of primordial Our results exhibit the existence of a critical ω∗ under black holes formed during the radiation era. In 1999, which the solution collapses to a black hole and above [22] performed simulations that showed that universality which it dilutes into the background. Concerning the holds also in the cosmological case, when considering a collapse case, we obtain a scaling law similar to (1), non empty backgroung universe. However, [23] showed with |ω − ω∗| as control parameter, in the Minkowski in 2002, in a similar case, that some families of initial and the de Sitter cosmologies. In a full matter Einstein- conditions admit a lower bound for the mass of a formed de Sitter universe, we observe a breaking of the univer- black hole : the scaling law (1) did not work for val- sality when using conformally flat initial conditions in ues of k very close to the critical solution but the mass the sub-horizon regime. But when using a super-horizon −4 seemed rather to stabilise towards 10 units of horizon fluctuation with exclusively growing modes (the long- mass. The authors explained that shocks, which are nu- wavelength solution), the universal scaling law appears merically challenging difficulties, are present when taking to be true. very small |k −k∗| and this should be the reason why [22] This work is interesting from GR and mathematical could not observe this phenomenon. points of view, but it has also important cosmological Universality in gravitation collapse is thus a widely motivations. Indeed, the large scale structure formation studied concept since the development of numerical rel- mechanism is still not completely understood and our ativity. Scaling laws similar to (1) have been searched in work could be a starting point for the study of spherically many other situations such as charged black hole mass, symmetric fluctuations evolution at several cosmological angular momentum, coupled scalar field, higher dimen- epochs. It could extend from primordial black holes for- sions,... (see [18]). In this article, we try to answer to mation in the radiation era to long term evolutions going the open question of the universality with respect to the through the equivalence radiation-dust epoch by consid- matter type, in the case of a barotropic perfect fluid with ering two-fluids simulations. linear equation of state p = ωe. Indeed, at fixed initial We organize the paper as following. Section II presents data, varying the parameter ω will intuitively divide the the BSSN formalism and all the equations that we use. solutions space into collapsing and non collapsing solu- The integration method, the gauge conditions and the ∗ tions, separated by a critical solution ω which should choice of the initial data are described in section III. The inevitably be the corresponding Evans-Coleman CSS so- section IV is devoted to the question of the initial condi- ∗ lution. The idea is thus to see if |ω−ω | can be considered tions. The section V contains all the numerical results, as a control parameter, as well as |k −k∗| was in previous including the code validation. We give our conclusions cases, and if a similar scaling law is verified. and perspectives in the last section of the article. To perform all this, we are following the works made in [24] and [25], within the framework of numerical rela- tivity. We use the BSSN formalism of GR, in spherical symmetry, conjointly with the Valencia formulation for II. EVOLUTION EQUATIONS the hydrodynamics [26]. Many numerical simulations in spherical symmetry use formalisms specially adapted to We give here a summary of the formalism we used this kind of symmetry, such as the Misner-Sharp formal- to solve numerically Einstein and hydrodynamical equa- ism (see [27] or [28] for a presentation of this formalism). tions. In all what follows, we work in natural units in A drawback of these formalisms is that they are written which G = c = 1. The time scale tscale, in s, is fixed in comoving gauges. Because we intent to study in the through the comparison of the experimental value of the exp future the spherical collapse of several fluids with rela- Hubble factor measured today H0 ∼ 70km/s/Mpc with tive velocities, comoving gauges were not suitable at the the adjustable parameter H0 in arbitrary units, chosen moment. The polar-areal gauge is often used too, such for numerical reasons : as in [29], but is difficult to use in a cosmological context since it does not converge to the FLRW metric as r → ∞. H0 tscale = exp . This is why we chose the BSSN formalism. When spher- H0 3

The length scale lscale, in m, and mass scale mscale, in kg, density (or the particle number density) ρ, the specific are thus computed through enthalpy h, the pressure p and the fluid 4-velocity uµ :

lscale = ctscale T µν = ρhuµuν + pgµν . (9) c3 m = t . scale G scale Spherical symmetry imposes that the only independent r r θ quantities are E, j , Sa := Sr and Sb := Sθ . A. BSSN formalism in spherical symmetry Following [32] and [24], the evolution equations for all the dynamical variables are : We follow what was made in [24] and because of spher- ical symmetry, we write the metric line element as ∂taˆ = −2αaAˆ a, (10)   2 2 2 2 4 2 2 ˆ 2 2 ˆ ˆ ds = −(α −β )dt +2βdrdt+ψ a (t) adrˆ + br dΩ ∂tb = −2αbAb, (11) (2) 1 1 a˙ ∂ ψ = − αψK − ψ, (12) where α(t, r) is the lapse, β(t, r) the radial component of t 6 2 a ˆ   the shift vector,a ˆ and b are the non-zero components of 2 2 2 1 2 ∂tK = −∇ α + α A + 2A + K the diagonal conformal√ spatial 3-metric. The conformal a b 3 factor is thus ψ a and we have factored out the cosmo- +4πα (E + S + 2S ) , (13) logical scale factor a(t) which follows its own dynamics a b   (3) ! ruled by background dynamical equations. The BSSN r 1 2 (3) r 1 ∂tAa = − ∇ ∇rα − ∇ α + α Rr − R formalism ensures that det (ˆγµν ) = 1, whereγ ˆµν is the 3 3 ˆ2 conformal 3-metric, which translates toa ˆb = 1. We 16π have split the extrinsic curvature into its trace K and its +αKA − α (S − S ) , (14) a 3 a b conformally trace-free part Aˆij : r 2 ∂t∆ˆ = − (Aa∂rα + α∂rAa) 1 4 2 aˆ Kij = γijK + ψ a Aˆij, (3)   3 r 2 +2α Aa∆ˆ − (Aa − Ab) rˆb where γij is the spatial 3-metric. Spherical symmetry ˆ ξαh 2 ∂rψ impose that Aij has only two non-zero components Aa := + ∂rAa − ∂rK + 6Aa r θ aˆ 3 ψ Aˆ and Ab := Aˆ . Since Aˆij is traceless, we have that r θ ˆ! Aa + 2Ab = 0. 2 ∂rb i + (Aa − Ab) + − 8πjr , (15) The particularity of the BSSN scheme lies in the ad- r ˆb dition of the auxiliary 3-vector ∆ˆ i, which corresponds to the conformal connection : together with the evolution of the scale factor a(t) throug ˆ i jk ˆi ij ∆ :=γ ˆ Γjk = −∂jγˆ . (4) Friedmann equation. We ensure strong hyperbolicity of the equations by setting ξ = 2 (see [33] ). In spherical symmetry, the only non-zero component of In this formalism, the Hamiltonian and momentum this vector is (see [32]): constraint equations read " ˆ  # r 1 ∂raˆ ∂rb 2 aˆ ∆ˆ = − − 1 − . (5) 2 aˆ 2ˆa ˆb r ˆb H ≡ (3)R − A2 + 2A2 + K2 − 16πE = 0, (16) a b 3 As in [24], we restrict to the zero shift case, β = 0, for r 2 ∂rψ M ≡ ∂rAa − ∂rK + 6Aa simplicity. 3 ψ The energy source terms measured by an Eulerian ob- ˆ! 2 ∂rb server are expressed by the projections of the energy- + (Aa − Ab) + − 8πjr = 0. (17) r ˆ momentum tensor T µν : b

µν E = nµnν T , (6) Those two equations are used to monitor the reliability µν ji = −γiµnν T , (7) and the stability of the method, the so-called free evolu- µν Sij = γiµγjν T , (8) tion scheme. The behaviours of the variables near the origin satisfy where nµ = (−α, 0, 0, 0) is the four-vector field orthog- µν the following parity conditions to ensure regularity of the onal to the spatial hypersurfaces. The tensor T of a 1 terms in (see [32]) : perfect fluid can be written in function of the rest-mass r 4

where the fluxes Fr are

2  r  α ∼ α0 + O(r ), (18) Dv r √ r 2 F = −g Srv + p , (32) aˆ ∼ aˆ0 + O(r ), (19)   τvr + pvr ˆ ˆ 2 b ∼ b0 + O(r ), (20) 0 2 Aa ∼ Aa + O(r ), (21) and the sources are[54] A ∼ A0 + O(r2), (22) b b  0  ∆ˆ r ∼ O(r2), (23) √ 1 S = −g  T µν ∂ g   2 r µν  ˆ 0 0 µν where α0,a ˆ0, b0, Aa and Ab are functions of time exclu- −T ∇µnν ˆ sively. Moreover, terms in (Aa − Ab)/r and (1 +a/ ˆ b)/r  0  must also disappear near the origin, which means that √ 1 = −g −αT 00∂ α + T rr∂ γ + T θθ∂ γ (33). the two following conditions, the flatness regularity con-  r 2 r rr r θθ 0r rr θθ dition, must occur : −T ∂rα + T Krr + 2T Kθθ

We recall that those expressions are exact only in the 2 0 0 Aa − Ab ∼ O(r ) ⇔ Aa = Ab , (24) case of spherical symmetry and vanishing shift (β = 0). ˆ 2 ˆ General equations can be found in [26] and [31]. aˆ − b ∼ O(r ) ⇔ aˆ0 = b0. (25) Using these equations in this form will be problematic Numerically, it is challenging to build a code that veri- in the case of a non-constant background metric. Indeed, fies simultaneously the parity and flatness regularity con- the asymptotic value of the vector U is not√ well defined ditions. Several articles discuss this issue (see [34], [35] in spherical coordinates because the term γ diverges and [36]). On our side, we do not need to implement such when r → +∞. To overcome this difficulty, we use the regularization conditions thanks to the kind of numerical reference metric approach presented in [38]. It consists code we use (see section III). rγ in taking as new variables U˜ = (D,S , τ), whereγ ˜ γ˜ r ij is a reference metric whose determinantγ ˜ is constant in B. Valencia formulation for relativistic time. In this case, the new fluxes are hydrodynamics with a reference metric  r  r Dv ˜ r γ r The evolution of the source terms can be written in F = α  Srv + p  , (34) γ˜ r r a conservative form by using the Valencia formulation τv + pv (see [26]) which ensures stability. To do this, we follow and the source terms are √[37] (and its notations) and we define the vector U = γ (D,S , τ) containing the conserved variables : r  0  rγ 1 D = ρW, (26) S˜ = α −αT 00∂ α + T rr∂ γ + T θθ∂ γ  γ˜  r 2 r rr r θθ S = ρhW 2v , (27) 0r rr θθ r r −T ∂rα + T Krr + 2T Kθθ τ = ρhW 2 − p − D. (28)  −F˜ r Γ˜k  D rrk where vr is the physical 3-velocity of the fluid for an  r  r k  γ  θ φ  + F˜ Γ˜ − Γ˜ + α p Γ˜ + Γ˜  .(35) Eulerian observer and W is the Lorentz factor :  Sr rr rk γ˜ θr φr  r ˜ r ˜k u −Fτ Γkr vr = , (29) αut The choice forγ ˜ is the flat metric in spherical polar 1 ij W = αut = √ , (30) 2 2 2 r coordinates :γ ˜ij = diag(1, r , r sin θ). Our final source 1 − vrv terms are thus, after evaluating the connection symbols, r with v in units of c.   We point out that it is generally not possible to re- 0 rγ 1 2 cover the primitive variables (v , h, p, ρ, . . . ) from the S˜ = α −αT 00∂ α + T rr∂ γ + T θθ∂ γ − F˜ r. r γ˜  r 2 r rr r θθ r conserved ones (D, Sr, τ) in an analytical way, except in −T 0r∂ α + T rrK + 2T θθK few particular cases. A root-finding procedure must be r rr θθ (36) used (see [37]). Note that this expression can be derived in a direct way The hydrodynamical equations ∇ T µν = 0, jointly with µ by simply developing the term the baryon number conservation, thus read   r r 2 ˜ r ˜ r 2 ˜ r ∂tU + ∂rF = S, (31) ∂rF = ∂r r sin θF = 2r sin θF + r sin θ∂rF 5 and inserting it in (31). where e = ρ(1 + ) = ρh − p is the energy density. This With that choice, our variables U˜ are well defined at equation is the limit   1, with ω = γ − 1, for flu- spatial infinity because ids which do not have a rest-mass density (such as ra- diation). The variable ρ is therefore not used in this rγ case. This equation of state has the advantage to give α = αψ6a3 → αa3 as r → ∞, γ˜ a simple (and analytical) formula to recover the primi- tive variables from the conserved ones (see Appendix A). where α is the asymptotic lapse. The value ω = 0 represents a pressure-less matter (dust) The last question that remains to be treated is to know 1 while the value ω = states for a radiation fluid (rel- what happens when we consider several matter fields. 3 The easiest way to proceed is to build a stress tensor 1 µν ativistic particles). Finally, ω < − is the condition T(k) for each matter field k and to sum them all to find 3 the total stress tensor : for the growth of the universe to be accelerated and the simplest way to achieve it is to consider a cosmological µν X µν constant in the Einstein equations, corresponding to a T = T(k). k constant value ω = −1. However, in this work we con- sider mostly universes not accelerated, only filled with The corresponding source terms can be summed in the matter fields with ω ∈ [0, 1]. same way : We here point out the fact that, when dealing with ho- mogeneous cosmological , the energy density X E = E(k), is usually denoted by the letter ρ in the literature. In our k case we prefer to use the letter e and use the letter ρ to X (k) denote the rest-mass density. jr = j , r Our code permits to have two different kinds of matter k with two different equations of state. The two fluids are X (k) Sa = Sa , considered as non coupled and thus are separately con- k served. For example, we can run a simulation with dust X (k) and radiation by choosing p = 0 and p = e /3. Sb = Sb . 1 2 2 k

Concerning the hydrodynamics equations (31), it de- III. IMPLEMENTATION pends on the adopted coupling between the fluids. In this work, we assume that there is no coupling between the To solve the hydrodynamical and BSSN equations, we fluids : each of them is conserved, independently from µν use the same method as in [24] (and first developped the others. We thus have ∇µT(k) = 0 for all k and one in [31]). The radial dimension is discretised by a uni- set of hydrodynamical equations for each matter field. formally cell-centered grid. A fourth-order finite differ- ence scheme is used to compute radial derivatives and we use fourth-order Kreiss-Oliger dissipation. A few virtual C. Equation of state points of negative radius are added to the grid to improve stability for the radial derivatives close to the origin by To close the system, we need an equation of state using parity conditions on the fields. p f(p, ρ, ) = 0, where  = h − 1 − is the specific in- We use the PIRK methods to solve the evolution equa- ρ tions. To achieve it, we split the set of equations in two ternal energy, which will describe what kind of fluid we parts : are using. If we want to simulate an ideal gas, the equa- ( tion of state will be of the form ∂ u = L (u, v), t 1 (40) ∂ v = L (u) + L (u, v). p = ρ (γ − 1) , (37) t 2 3 The variables u are first explicitly evolved and the re- where γ is the adiabatic index. For a polytropic fluid, sult is used to evolve v partially implicitly through the the equation will be operator L2. Since it is a second order PIRK method, p = KρΓ, (38) the evolution requires two steps which are described in details in [31]. In particular, if we denote by L1, L2 and where K is the polytropic constant and Γ is the poly- L3 the corresponding discrete operators of L1, L2 and tropic exponent. Those two cases are widely used in nu- L3 (the exact expressions for the splitting operators are merical relativity simulations. However, in cosmology we given in Appendix B.), the operators L1 and L3 are used often work with a linear barotropic equation of state : in an explicit way, while L2 contains the unstable terms and is treated in a partially implicit way. This method p = ωe, (39) has already been used in the frame of BSSN formalism 6 under asymptotically flatness assumption (see [39]). It needed are the following ones for the rest-mass density has also been applied for a dynamical cosmological back- and the energy density : ground in [24] and [25], but it was restricted to the case a˙ of dust matter (pressureless matter) and scalar field, and ∂tρ = −3 ρ = αKρ , k a k k so it did not really include hydrodynamics. The vari- a˙ ables which are first explicitly evolved (those contained ∂tek = −3 (ek + pk) = αK(ek + pk), in the vector u) are the hydrodynamical conserved vari- a 3 a˙ ables, the cosmological scale factor a, the lapse α, the where K = − is the trace of the homogeneous extrin- elements of the conformal 3-metrica ˆ and ˆb as well as ψ. α a sic curvature. Indeed, the other hydrodynamical vari- Their updated values are subsequently used to evolve K, ables can be recovered by using only the equation of state A and ∆ˆ r. a because the velocity v r is null and the Lorentz factor is Concerning the hydrodynamical equations, we first use k thus equal to 1. a monotonised central-difference (MC) slope limiter (see [40]) to approximate the left and right states of the prim- itive variables at each cell. Secondly we solve the equa- 2. Boundary conditions tions with a HLLE incomplete Riemann solver (from Harten, Lax, van Leer [41] and Einfeldt [42]). Finally, we use a root-finding procedure (Newton-Raphson) to The spatial domain is of the form r ∈ [0, rspan], where recover the primitive variables from the conserved ones 0 corresponds to the origin and rspan to the outer bound- if the equation of state is different than (39). ary. We use a cell-centered discretization to avoid cal- culations at the exact origin in case of singularities. At the origin, we impose, following spherical symmetry, the A. Gauge conditions inhomogeneous variables to have the correct parity for a regular solution thanks to a few virtual points of negative 1. Background evolution radius we added to the grid. At the outer boundary, we use a Sommerfeld (radiative) boundary conditions (see [43]) : we impose the variables to behave like outward As we said, we consider models in which space-time is travelling waves when r is near rspan. This means that, not asymptotically constant but rather looks like a ho- at a few outermost points of the computational grid, any mogeneous FLRW Universe, without curvature, at large field f(t, r) must verify radii. Note that in all what follows, an overline is used v  to indicate the background value of the quantity. The ∂tf = ∂tf − v∂rf − f − f , (45) line-element is given by the FLRW metric, with possibly r a dynamical lapse : where v is the characteristic velocity of the field. Note that v is computed by examining the dynamical equation of each field and is the for most of it. Only 2 2 2 2 2 2 2 ds = −α (t)dt + a (t) dr + r dΩ , (41) the lapse α and the variable ∆ˆ r admit a characteristic velocity different from it. For the lapse, it depends on where we consider that there is no curvature. The evolu- √ the slicing that is used while it is 2 for ∆ˆ r. Such a tion of this metric is ruled by the well known Friedmann condition prevents any signal to be reflected by the outer equations boundary. The asymptotic values of each variables are 1 a˙ 2 8π the homogeneous ones given by the background evolution = e, (42) : α2 a 3 1 a¨ a˙ α˙ 4π α(t, r) → α(t), (46) − = − (e + 3p) , (43) ˆ α2 a a α3 6 aˆ(t, r), b(t, r), ψ(t, r) → 1, (47) 3 a˙ where e and p are the homogeneous background energy K(t, r) → K(t) = − , (48) density and pressure. These two quantities are composed α a ˆ r with the contribution of different kinds of energy (matter Aa(t, r),Ab(t, r), ∆ (t, r) → 0, (49)

(in general several species) and possibly a cosmological ρk(t, r) → ρk(t), (50) constant Λ in our case) : ek(t, r) → ek(t), (51)  X r e = e + e , vk (t, r) → 0. (52)  k Λ  k X (44) p = p + p .  k Λ 3. Slicing conditions  k For the background, the hydrodynamical equations are There are lots of different slicing conditions in the liter- simplified and the only evolution equations that are ature (see for example [44], [27] or [28]). We implemented 7

G M the Bona-Masso slicing ([45]) for the local dynamics and : . Similarly, we define the compaction function as the geodesic slicing (constant unity lapse) for the back- c2 R ground : the mass excess inside the sphere with areal radius R :   2pˆ ∂ α = −α2f(α) K − K , (53) 2 MK (t, r) − MK (t, ψ br) t C(t, r) = , (55) α = 1. (54) R where the factor 2 is simply a question of convention to We differ a bit from what was made in [24] and [25]. follow the definition of [49]. This quantity is useful to They considered a slicing condition that did not converge define the size of the fluctuation. The radius r (t) is de- to the geodesic one at spatial infinity. Our slicing condi- m fined, in hand, as the coordinate r where the compaction tions are thus not exactly the same, we added the −K in function reaches its maximal value. With this point, we our equation to ensure that α → 1 as r → ∞. This allows can define the mass of the fluctuation as identifying the time coordinate to the cosmological syn- chronous time for physical interpretations. Such a slicing Mm(t) := MK (t, rm(t)) (56) condition gives a characteristic velocity for the lapse of p f(α = 1) and its compactness as αpf(α)γrr (see [33]), which is equal to at a spatial infinity. This quantity must then be taken in the Cm(t) := C(t, rm(t)) = max C(t, r). (57) Sommerfeld conditions of (45). r>0 1 Choosing f ≤ implies that the coordinate speed The last quantity we need before using the code is the 3 amplitude of the fluctuation. The standard way in cos- of light remains finite (see [46] ). Thus, this condition mology is to take the central value of the energy density ensures to keep the stability of the scheme though not contrast. But in fact, there is no reason to consider only mandatory. this particular value because this quantity is not neces- The simplest choice f = 0 is the geodesic slicing α = 1. sarily representative of the full behaviour of the fluctua- Combined with a zero shift β = 0 gives what is called tion, especially when pressure enters into consideration. the synchronous gauge. Although this is not the best Indeed, the energy-density contrast radial profile changes choice in term of stability - it easily generates coordinate when pressure increases. This is why we use an average singularities (see [44])- we chose this one in some of our energy density contrast defined by simulations because of its simplicity. To perform the simulation of black holes formation, we R R 4πδR02dR0 2 δ(t, r) = 0 . (58) use the 1 + log slicing defined by f(α) = . This seemed R R 4πR02dR0 α 0 to be to most efficient Bona-Masso slicing in this context. The mean energy-density contrast of the fluctuation, i.e. its amplitude, is defined as this quantity evaluated at the radius of the fluctuation : B. Used quantities

δm(t) = δ(t, rm(t)). (59) In , deriving a well-posed definition of mass is a difficult challenge since it should be gauge in- The pertinence of this definition can be pointed out variant and, if possible, time invariant. Moreover, several through the following relations obtained in the long- notions of mass exist in general relativity. For asymptot- wavelength approximation : ically Minskowski spacetimes, there still exists the pos- δ (t) ' 3δ(t, r (t)) (60) sibility to compute the total energy on a single slice Σt. m m 2 The ADM mass (see [28]) is based on this principle and Cm(t) ' δm(t)(H(t)R(t, rm(t))) , (61) gives a time independent quantity. Another mass notion that is often used in static spacetimes is the Komar mass where the first approximation has been derived in [49] (see [44]). But since we are working in an asymptotically and the second in [47]. These relations show that the Friedmann universe, we cannot use those definitions with particular radius rm(t) contains an important part of the no change. Indeed, the total mass is not finite and we information characterizing the fluctuation whatever its need to truncate the computation to keep only the mass shape. inside the fluctuation. To achieve that, we follow [13] and To conclude, the quantities we will focus on are the [47] and use the Kodama mass, which was first define in following ones: [48]. The definition et expression of this mass is given in details in Appendix C. Now that we have defined our - the central energy density contrast δ0(t); mass notion, we are ready to use it to define a compact- - the mean energy density contrast δm(t); ness notion. Usually, the compactness of an object is a dimensionless quantity defined as its ratio mass to radius - the compactness Cm(t); 8

- the radius rm(t) and the corresponding areal radius rest-mass density quantity ρ. Only two last quantities Rm(t); thus remain to be specified : the energy density and the initial curvature, i.e. the conformal factor. Both are - the mass Mm(t); linked by the Hamiltonian constraint : - the central value of the lapse α (t) when a non 0  2  3 geodesic slicing is used. −a−2ψ−5 ∂2ψ + ∂ ψ + H2 = 2πE(t = 0, r). (67) r r r 4 i So, specifying one determines the other. Intuitively, IV. INITIAL CONDITIONS the physicist prefers to introduce a specific energy- density profile to model a situation. This is what was There are two ways of setting the initial conditions : done in [24], [25] and [13]. The initial conformal factor either we choose well determined values for almost all ψ(t = 0, r) is then found by solving numerically (67) as the variables and derive the last ones with the constraint a boundary value problem with equations (16) and (17), or we choose initial conditions following a good approximation of the solution modelling ∂rψ → 0, for r → 0; (68) the situation at early times. The first method has the Cψ advantage to start with, theoretically, an exact solution ψ → 1 + , for r → ∞, (69) 2r of general relativity since the constraint equations are verified at initial time. But the initial data generated where Cψ is a constant adjusted such that in this way can be viewed as quite artificial and could suffer from transient effects at early times before reach- Cψ ∂rψ → − , for r → ∞. (70) ing a more realistic evolution. The second method has 2r2 opposite properties. It generates a more realistic initial This method is used successfully in section V A and situation from the physical point of view but this is gen- in all the following chapter. We specify the initial en- erally only an approximation of the exact solution and ergy density in terms of the energy density contrast then needs some iterations for the constraints to reach a i ek(t = 0, r) more acceptable level of accuracy. We start by exhibiting δk(r) = − 1. We use a smooth top-hat pro- ek(t = 0) the first method in a simple but useful case. file (a logistic function) to simulate the evolution of an over-dense region of the universe :

A. Conformally flat initial conditions  i  r−rk 1 − tanh 2σ δi (r) = δi k , (71) k k  ri  This idea here is to assume spatial homogeneity on 1 + tanh k each metric variables except the conformal factor and 2σk thus to equal them to their corresponding background i i where δk, rk and σk are positive parameters designing values : the shape of the profile.

α(t = 0, r) = 1, (62) B. The long-wavelength approximation aˆ(t = 0, r) = ˆb(t = 0) = 1, (63)

K(t = 0, r) = Ki = −3Hi, (64) To generate more realistic initial conditions, we can Aa(t = 0, r) = Ab(t = 0, r) = 0, (65) use the long-wavelength approach (also called gradient expansion), as described in [50], [47] and [49]. This ap- ∆ˆ r(t = 0, r) = 0, (66) proach resembles the cosmological perturbation theory where Ki is the initial background curvature and Hi the but, instead of developing the equations in powers of the initial Hubble factor. inhomogeneities as in the lattice, the long-wavelength Concerning the matter source terms, the Momentum scheme expands the solution in the spatial gradient of constraint (17) reduces to jr(t = 0, r) = 0. This imposes these perturbations. Concretely, we focus on some fixed a null spatial 3−velocity : time, multiply each spatial gradient ∂i by a fictitious pa- rameter   1 and we expand the equations in a power r vk(t = 0, r) = 0. series of . We keep terms up to first-order in  and then set  = 1. To compare, in the perturbative approach the Note that in the case of several fluids, the condition parameter  would multiply the perturbation instead of jr(t = 0, r) = 0 is weaker than vr(t = 0, r) = 0 for k ∂ . all k. But other possibilities are much more complicated i The parameter  is conveniently identified as the ratio and, we think, naturally improbable. Since we will work with the cosmological equation of R (t)  := H , (72) state p = ωe, we recall that we do not have to use the L 9

1 where R (t) = is the Hubble radius, the only ge- we have the following expressions for pa, pb and F : H H(t) ometrical scale in the homogeneous universe, and L is 2  1  p = − e−2ζ ζ00 + ζ0 − ζ0 , (87) the (comoving) length scale of the perturbation. This a 3 r approach reproduces the results of linear perturbation    1 −2ζ 00 0 1 0 theory but can also consider non linear perturbations of pb = e ζ + ζ − ζ , (88) 3 r the curvature if the universe is sufficiently smooth for 2  2 ζ0  scales greater than L (see [50] and [49]). F = − e−2ζ ζ00 + ζ0 + . (89) The complete derivation of the solution in a general 3 r 2 Bona-Masso gauge (53) is presented in the Appendix D. In terms of BSSN variables in spherical symmetry, the Moreover, if we choose the Gaussian profile for ζ, i.e. long-wavelength solution is given by  r2  ζ(r) = A exp − 2 , (90) 2  2! 2∆ 1 v∗ 1 ψ = Ψ 1 − 2 1 F 6(1 + ω) v∗ − 2v∗ aH these expressions become +O(4), (73) 4ζ  r2  p = e−2ζ 1 − (1 − ζ) , (91)  2 a 3∆2 2∆2 2 1 4 Aa = paH + O( ), (74)  2  3ω + 5 aH 2ζ −2ζ r pb = − e 1 − (1 − ζ) , (92)  2 3∆2 2∆2 2 1 4 Ab = pbH + O( ), (75) 2ζ  r2  ζ  3ω + 5 aH F = e−2ζ 1 − 1 + . (93) 2 ∆2 3∆2 2 4  1  aˆ = 1 − p + O(4), (76) (3ω + 5)(3ω + 1) a aH As a confirmation of the development, the particular  2 case of the geodesic slicing with f(α) = 0 gives a solution ˆ 4 1 4 b = 1 − pb + O( ), (77) that is identical to the one developed in [47]. This solu- (3ω + 5)(3ω + 1) aH tion will give realistic initial conditions for the spherical v2  1 2 collapse if the length-scale of the inhomogeneity is greater δ = ∗ F + O(4), (78) 2 1 than the Hubble radius. v∗ − 2v∗ aH 1  2! v∗ 1 4 K = K 1 + 2 1 F + O( ), (79) v∗ − 2v∗ aH V. NUMERICAL RESULTS v3  1 2 ∗ 4 We start now the section which presents our all our α = 1 + 2 1 F + O( ), (80) v∗ − 2v∗ aH numerical results. To avoid the reader to be lost, we give 2 3  3 2 ωv∗ + (1 + ω)v∗ 1 here a little explanation of the process presented in this vr = − · 2 1 ∂iF a section. We first validate the code in section V A. Then, 3ω + 5 (1 + ω)(v∗ − 2v∗) aH 5 we present, in section V B, the two typical behaviours +O( ). (81) that we observed in our simulations : the collapse and where dilution scenarios. The next step is the choice of a gauge that permits to simulate black holes formations. This is  1  2  v∗ (1 + 3ω) done in sectionV C where simulations of sub and super- 2 v∗ = −3(1 + ω)(1 + 3ω + 3f(1)) . (82) critical solutions are performed. Finally, the universality 3 of the phenomenon is investigated in section V D with v∗ 3(1 + 3ω)f(1) this specific gauge. and we used the intermediate variables pa, pb and F given by A. Code validation     1 4 2 1 4 2 pa := − ∂r Ψ − ∂rΨ + (∂rΨ) ,(83) Ψ4 3Ψ r Ψ2 To achieve this validation, we performed a simulation     1 2 2 1 2 2 with two species of matter which have linear equations of pb := ∂r Ψ − ∂rΨ − (∂rΨ) , (84) Ψ4 3Ψ r Ψ2 state p1 = 0.1e1 and p2 = 0. We chose the geodesic slic- ing and the initial data are fixed by the conformally flat 4 ∆Ψ method discussed in section IV A. Note that we do not F := − 5 . (85) 3 Ψ validate our code with the long-wavelength approxima- If we write tion as initial conditions since we need initial conditions perfectly correct (not an approximation) from the gen- ζ Ψ := e 2 , (86) eral relativity point of view. The initial profiles for the 10

Figure 1: Hamiltonian constraint at t = 25 for simulation of the Figure 2: Approximation of the L2-norm of the Hamiltonian con- evolution of a smooth inhomogeneity in the density profile, with straint for simulation of the evolution of a smooth inhomogeneity three resolutions : ∆r = 0.1, ∆r = 0.05, and ∆r = 0.025. Curve in the density profile, with three resolutions : ∆r = 0.1, ∆r = 0.05, for ∆r = 0.05 has been multiplied by 4 and curve for ∆r = 0.025 and ∆r = 0.025. Curve for ∆r = 0.05 has been multiplied by 4 has been multiplied by 16 to exhibit the second order of convergence and curve for ∆r = 0.025 has been multiplied by 16 to exhibit the of the method. second order of convergence of the method.

ek tonian constraint with respect to time, in Fig. 2, we energy density contrasts δk = − 1 are of the form of ek see that we also have a second order rescaling (as in the the smooth top-hat functions given by eq. (71). We fix i i previous plot, curve for ∆r = 0.05 has been multiplied the positive parameters of these profiles to δ1 = δ2 = 1, i i by 4 and curve for ∆r = 0.025 has been multiplied by r1 = r2 = 10 and σ1 = σ2 = 1, while our spatial do- 16). The convergence of the method is thus at least sec- main is the interval [0, 500] (all in code units). We put ond order. Note that the late but steep increase at the as much quantity of matter 1 as of matter 2, which reads i i end of the simulation is due to the collapse and possi- Ω1 = Ω2 = 0.5. bly the reaching of a singularity. Indeed, we can see on The last quantities that remain to be fixed are the Fig. 3 that the total energy-density contrast (defined by initial scale factor ai, the initial Hubble factor Hi and the e1 + e2 δtot = − 1, which is different from δ1 + δ2) seems Hubble factor measured today H0 which will determine e1 + e2 the time scale, the mass scale and the length scale. For to diverge. However, note that this is not a significant our tests, we chose ai = 1, Hi = 0.03 and H0 = 0.001. evidence that a black hole is formed. We will discuss this The Courant-Friedrichs-Lewy factor (CFL) is set to 0.25, in the next section. indicating that the discretization step in time ∆t is linked We thus have validated our integration code since the to the spatial discretization ∆r through the relation ∆t = Hamiltonian constraint has the correct behaviour. We 0.25∆r. We tested the three resolutions ∆r = 0.1, ∆r = are allowed to trust our simulations and are ready to 0.05, and ∆r = 0.025. investigate in the following sections the non linear cos- We now present the results of these simulations. The mological spherical collapse with this tool. Hamiltonian constraint at t = 25 is shown on Fig. 1. The similarity in the shapes of the curves and the fact that it is rescaling with the resolution in the right order (curve B. Typical behaviours and dependence on the for ∆r = 0.05 has been multiplied by 4 and curve for equation of state ∆r = 0.025 has been multiplied by 16) show the stability of the method and at least a second-order convergence of Before looking for universality with respect to the the scheme. Of course, the error is maximal at the centre equation of state, we study the different behaviours of of coordinates and at the boundary between the inner a fluctuation of a single fluid and the influence of the and outer parts of the over-density. Terms in inverse equation of state on it. In all the runs of our code, we fix power laws of the radius are responsible for the larger the following parameters (except if especially mentioned) error at the center. For the overdensity boundary , it : Hi = 0.03, ai = 1 and H0 = 0.001. is the location where the gradients are maximal and it We perform simulations involving a single barotropic justifies these peaks in the error. fluid of matter p = ωe with an initial profile described Moreover, by inspecting the L2-norm of the Hamil- by the equation (71), which is, as already mentioned, 11

Figure 3: Central total energy-density contrast versus time. The divergence indicates a collapse. parametrized by three real numbers: the initial ampli- tude δi, the initial size ri of the fluctuation and the sharp- ness of the profile σ. This profile and the correspond- ing initial compaction function are shown in Fig. 4. A compaction function has usually this bell shape: starting from zero, growing to a peak and then decreasing to zero as asymptotic behaviour. The peak determines, as de- fined in section III B, the size of the fluctuation. We can see that it is nearly the same value as ri, which confirms the pertinence of this definition. Depending on the three initial parameters, δi, ri and ω (σ is fixed to 10Mpc in all what follows), we observe two different behaviours. The first one is the collapse while the second one is dilution. Figure 4: The upper graph shows the shape of an initial profile of the energy-density contrast. The second one gives the correspond- ing compaction function. The peak in the latter is in agreement 1. The collapse scenario with the size of the fluctuation seen in the above panel.

We show such an example of collapsing solution in Fig. 5, 6, 7 and 8 with the values ω = 0.01, δi = 0.5 and nesses, but this is sufficient for our purpose. Note that ri = 100Mpc. We see on Fig. 5 that the central and the relation (61) is verified all along the collapse, even if mean energy-density contrast are both diverging on the we are not in a subhorizon regime. The reason why this first plot. We also see that the relation (60) seems to relation and (60) are verified must probably due to the be correct with good accuracy since the curves of δm low value of the equation of state parameter ω. Recall and 3δ(rm) are nearly the same, although the conditions that these relations are perfectly verified in the comoving for the long-wavelength approximation are not verified in gauge. And the absence of pressure precisely generates this situation. The last curve represents δlin, the central a comoving gauge. Thus the low pressure explains why energy-density contrast computed with the linear per- these relations seem to hold in this case. The Fig. 7 turbation theory (see [51] for the basic equations). The shows that the radius of the fluctuation has an increas- agreement between this curve and δ0 := δ(r = 0) at ing phase, coherently with the decreasing phase of the early times is an additional indication of the validity of compactness, followed by a fast decreasing. The matter the code. is concentrating towards the center of the grid, which is Concerning the compactness (Fig. 6), it is first de- intuitively logical in the collapse scenario. The Fig. 8 creasing because of the background expansion. But then shows a decreasing mass. Although this can seem to be it grows until the end of the simulation, indicating a col- illogical, this is normal because the mass we used is an lapse. Our code is not able to follow it at higher compact- integral whose upper bound is rm, which is decreasing. 12

Figure 5: Evolution of the central energy-density contrast δ0, the Figure 7: Evolution of the radius of the fluctuation. The steps mean energy-density contrast δm and the central energy-density come from the spatial discretisation. The radius first increases contrast δlin computed with the linear perturbation theory in the with the background but then starts decreasing, indicating a con- collapse scenario. The last curve, 3δ(rm), shows the validity of the centration of the matter towards the center of the grid. formula (60), although we are not in a super-horizon regime. The full relativist δ0 and the approximate δlin are in adequation at early times, which is an additional validation of the code.

Figure 8: Evolution of the mass of the fluctuation, in solar mass units. The mass decreases because the integration upper bound is related to the radius of the fluctuation, which is collapsing to zero. Figure 6: Evolution of the compactness of the fluctuation. The 2 second curve represents the quantity δm(HR) and its adequation with Cm confirms the approximation (61), although we are not in a super-horizon regime. The compactness decreases first with the background expansion but then increases more and more rapidly 2. The dilution scenario with the collapse. The code is not able to follow it until the black hole formation at Cm = 1.. We give now an example of diluting solution in Fig. 9, 10 and 11 with the values ω = 0.1, δi = 0.1 and The steps visible in these two graphs are simply due to ri = 100Mpc. First notice that the two correspondences the spatial discretization of the grid. Indeed, the radius (60) and (61) are no more exactly verified because we rm, i.e. the location of the maximal compaction, can involved more pressure and the long-wavelength ap- only be taken among the spatial values given by the dis- proximation does not hold. Then, on the first plot, the cretization, explaining the discontinuous bumps revealing central energy-density contrast has an oscillations phase a change in this value. before decreasing to zero in a power law of a. This is 13

Figure 9: Evolution of the central energy-density contrast δ0, the Figure 11: Evolution of the radius of the fluctuation. It is increas- mean energy-density contrast δm and the central energy-density ing in a powerlaw of the background scale factor, indicating that contrast δlin computed with the linear perturbation theory in the the fluctuation follows the cosmological expansion. dilution scenario. The last curve, 3δ(rm), shows the inaccuracy of the formula (60) in this case. The full relativist δ0 and the ap- proximate δlin are in adequation at early times. The full relativistic solution is shown to be in adequation with the linear solution. This confirms the validity of the numerical computation. and make it disappear.

C. The 1 + log slicing to simulate black holes formation

We now test another slicing condition, the 1 + log slic- ing presented in section III A 3, to try to follow the col- lapse until the formation of a black hole. With this slic- ing, we give here an example of a sub and a super-critical solution obtained in that gauge. The central value of the lapse is shown in Fig. 12. The sub-critical solution is a collapsing solution and we can see the formation of a black hole since the lapse is collaps- ing to zero. This freezing of time at the centre permits the code to avoid the central singularity and to pursue the calculation outside the horizon after its formation. On the contrary, the other solution exhibits a lapse re- growing away from zero, indicating that the overdensity Figure 10: Evolution of the compactness of the fluctuation. The 2 is diluting. At early times, both solutions follow nearly second curve represents the quantity δm(HR) and is no more in the same curve before choosing between a collapsing or a adequation with Cm. The compactness decreases in a powerlaw of the background scale factor, which means that the fluctuation diluting solution. This is typically what occurs in critical follows its expansion. phenomena : near-critical solution all follow the critical (self-similar) solution until falling into a black hole solu- tion or a diluting one. The time before leaving this curve in agreement with the linear perturbation theory, which depends on the closeness to the critical parameter ω∗. gives another validation of the numerical computations. The compactnesses are presented in Fig. 13. The The mean energy-density is also decreasing but without collapsing solution shows a compactness growing higher oscillations. Fig. 10 and 11 show, through clear power than unity, the black hole limit. But the final compact- laws in a, that the fluctuation is at late time completely ness of the object seems to stabilise to a value near 1.4. diluted in the background and only follows the dynamics On the other hand, in the diluting solution, the compact- of the latter. The background expansion and the internal ness decreases to zero, indicating that the over-density is pressure are too strong for the fluctuation to collapse progressively disappearing. 14

Figure 12: Central value of the laps in a sub and a super-critical Figure 14: Evolution of the mass in a sub and a super-critical solution. The sub-critical solution leads to the formation of a black solution. In the collapsing solution, the mass is decreasing until hole as the collapse of the lapse indicates it. The super-critical the black hole formation. After that, it regrows, showing that solution shows a lapse regrowing indicating a dilution. the formed black hole attracts matter. The other solution shows equally a growing mass because of the radius of the fluctuation grows with the background expansion.

Figure 13: Evolution of the compactness in a sub and a super- critical solution. The compactness reaches values greater than unity for the sub-critical solution, revealing the formation of a black hole. Its final compactness seems to be around 1.4. The diluting super-critical solution shows a decreasing compactness. Figure 15: Evolution of the areal radius in a sub and a super- critical solution. The collapsing solution shows a size that is de- creasing until the formation of a black hole and after that is re- growing. Concerning the mass, we see in Fig. 14 that it is de- creasing until the formation of the black hole. Then, once it has formed, the black hole grows and accretes matter. The diluting solution shows a late-time evolution following the cosmological expansion and thus a late-time ”Which value for the mass of the formed black hole growing mass because of the growing of rm. should we take?”. As we saw, it is not stabilizing to The shape of the areal radius, in Fig. 15, equally shows a constant value but is growing after the black hole for- that the size of the black hole regrows once it has formed, mation. Thus, the most reasonable choice is to take the due to the matter accretion, and that, in the diluting value of the mass when the compactness is equal to one, solution, it is always growing. which corresponds almost to the minimum in the curves With all these information, the question is thus of the mass and of the areal radius. 15

D. Universality in the critical phenomenon

We now want to see if the spherical collapse is univer- sal with respect to matter species. For this reason, we must check if the relation (1) is verified if we take for the parameter k the equation of state parameter ω (although this parameter does not represent exactly what is called strictly speaking the initial conditions), that is to say :

M ∝ |ω − ω∗|γ (94) with γ a constant independent of ω. We test this hypoth- esis in three different cosmologies : the empty -time, the empty de Sitter space-time and the full of matter Einstein-de Sitter space-time.

1. The Minkowski case

Figure 16: Mass of the formed black hole in a Minkowski back- This case is, by far, the easiest one. Our code was ground computed with the 1 + log slicing. The power law is clearly not built to deal with an empty background but the visible, proving universality in this case. only differences consist in the scales and the initial con- ditions. The time, length and mass scales are, here, GM GM given by t = (in s), l = (in m) and scale c3 scale c2 mscale = M (in kg), while the initial profile is based on the energy-density profile instead of the energy-density contrast profile because em = 0 :

 i  r−rm 1 − tanh 2 ei (r) = ei , (95) m m  i  rm 1 + tanh 2

i i where em is the initial amplitude of the object and rm its initial radius. We work in code units and take as initial i −5 15 conditions em = 10 (which corresponds to 6.18 × 10 kg ) and ri = 20 (which corresponds to 2.95 × 104 m), m3 m with an initial compactness of 0.048. We use the 1 + log slicing to observe the complete formation of black holes. All this gives us as critical ω∗ the value of 0.0094 and the evolution of the mass of the black hole, that we follow Figure 17: Mass of the formed black hole in a de Sitter background until its formation, with respect to |ω − ω∗| is shown computed with the 1 + log slicing. The power law is clearly visible, proving universality in this case. in Fig. 16. In this plot, we observe that all points lie nearly perfectly along a straight line, indicating a power law and thus universality. This is in agreement with [17] With this, the critical solution appears to be around and generalises universality to one particular 1-parameter ω∗ ' 0.0128. The mass of the formed black holes, with family of matter species, in a Minkowski background. respect to |ω − ω∗| is shown in Fig. 17. We observe that all the points are along a straight line, revealing the 2. The de Sitter case universal scaling law. As well as in the Minkowski case, we thus proved universality in the de Sitter case which is relevant from the cosmological point of view. Our second test will be the case of an empty space with a positive cosmological constant Λ. We use the same i −5 energy profile as in the previous test with em = 10 and i −5 3. The Einstein-de Sitter case rm = 20. The initial Hubble factor is set to Hi = 3×10 and ΩΛ = 1, in such a way that the value of Λ is fixed by Hi. The 1 + log slicing is used to follow the black holes The Einstein-de Sitter case is more complicated be- formations. cause of the presence of matter at the outer boundary. 16

Figure 18: Mass, by units of horizon masses, of the formed black Figure 19: Mass, by units of horizon masses, of the formed black hole computed with the 1 + log slicing. No scaling law is observed, hole computed with the 1 + log slicing for two families of profiles. ∗ ∗ indicating that values closer to ω = 0.058 should be computed to Black dots represent family 1, with ω1 ' 0.334, while green ones ∗ look for universality. represent family 2, with ω2 ' 0.434. In both cases, dots seem to be globally on the same line, indicating a power-law and, thus, that the universality relation is verified. The slope of the line depends This non zero asymptotic value for the energy-density on the family of profiles. renders the code less stable than in the two previous cases. However, this code is able to follow the forma- tion of a black holes even in this case. We will just not tion starting in the super-horizon regime that admits only be able to go as close to the critical solution as we did growing modes when entering the horizon. To achieve for the Minkowski and the de Sitter cases. Moreover, the that, we use initial conditions computed from the long- universal scaling law differs a bit from these two cases wavelength approximation presented in section IV B, still because it is not the mass of the black hole that is rescal- in the 1 + log slicing. The parameter ∆ of profile (90) M is chosen such that the initial areal radius of the fluctua- ing but the ratio where MH is the horizon mass −1 MH tion is three-times the Hubble radius RH = H . In such computed at the time when the areal radius of the fluc- a way, all the decreasing modes have disappeared when tuation is equal to the Hubble radius. We thus want to the size of the inhomogeneity coincides with the cosmo- prove that logical horizon. The second parameter, A, denotes the amplitude of the overdensity and will thus also influence M ∝ |ω − ω∗|γ (96) the value of the critical equation of state parameter ω∗. MH ∗ We perform two sets of simulation, one with ω1 ' 0.337 ∗ with γ a constant independent of ω. and the other with ω2 ' 0.434. To achieve that, we first test the conformally flat initial The masses of the formed black holes, rescaled by the conditions of Sec. IV A with the smooth top-hat profile horizon mass MH (the mass inside the horizon computed i (71). Values taken for the profile are δm = 0.5, ri = at the time the fluctuation enters the Hubble radius), 100Mpc and σ = 10Mpc and gives ω∗ = 0.058 as a critical are shown in Fig. 19. It is clearly visible, in both cases, value. that the plotted quantity is rescaling in a power-law of The mass values computed in this way are reported |ω − ω∗|, although our code does not permit to obtain on Fig. 18. Unfortunately, the results are not in favour perfect lines and values nearer the critical solution. This of universality : no scaling law is visible and, worse, no is thus a serious indication in favour of universality. Note decreasing of the mass is seen when considering the most that the critical exponent γ is shown to vary as a function left points in the graph. We are thus in a similar case of the amplitude of the profiles. to what is described in [23]. In [52], it is explained that We can draw two major conclusions from this. First, this behaviour appears because taking non linear initial our simulations with sub-horizon and conformally flat ini- profiles (with too large initial amplitude δi) whose length- tial conditions, though interesting and relevant, revealed scale are smaller than the cosmological horizon generates a breakdown of universality close to ω∗. Secondly, the shocks in near critical solutions because of the presence universal scaling law seems to be verified when using of decreasing modes. We will thus need to test super- super-horizon initial conditions with exclusively growing horizon fluctuations to try and avoid this behaviour. modes. This is perfectly in agreement with [23], [52] and We thus follow the advises of [52] and take a fluctua- [53]. They worked by fixing the equation of state and 17 varying the matter profile. On our side, we showed uni- is that we encountered a similar behaviour as in [23] who versality with respect to the matter type by varying the saw a breaking of universality near the critical solution equation of state parameter and this is the major orig- because of the use of conformally flat and sub-horizon inal result of this work. It should however be checked initial conditions that possess decreasing modes. To by- with a more efficient code, equipped with an adaptive pass this problem, we used super-horizon initial fluctua- mesh-refinement specially designed for spherical symmet- tions with exclusively growing modes. We thus derived, ric space-times like the one used by [53], to be proven in any Bona-Masso slicing given by eq. (53), the solution definitely. of the equations in the long-wavelength approximation. So, with all these simulations, using different gauges This original development, which is interesting in itself and backgrounds, we can be confident with our results because of its similitude to the same solution in [47] in and conclude that the spherical collapse is fully universal other gauges, allowed us to evolve more realistic initial (with respect to the equation of state) in a Minkowski conditions. With it, we could finally observe the uni- background and partially universal in a full matter back- versality scaling law we were looking for. The result is ground. We end by saying that we guess that the critical not perfect since we could not approach the critical solu- exponent of the power law must depend on the initial tion as close as we wished, an adaptive mesh refinement profile and its value, in itself, should thus be less funda- should be used to perform this, but it gives more than mental than those found by varying the initial conditions an indication that universality is true in this case. for a fixed matter specie. In conclusion, we have shown, among others, that the spherical collapse of a barotropic fluid p = ωe is universal with respect to the value of the equation of state param- VI. CONCLUSION eter. These results are important because it is a gener- alisation of numerous previous works: the gravitational In this work we have upgraded the BSSN code used spherical collapse appears to happen in a very general in [24] and [25] to study the spherical collapse of pres- way that depends little on the initial curvature profile sured matter thanks to the addition of a HLLE incom- and, we know it now thanks to our own work, the types plete Riemann solver for the relativistic hydrodynamics. of matter (through the equation of state parameter ω). This code is now able to deal with several matter species, Our treatment of the problem enforces the importance of thanks to the use of a non comoving gauge, in a general the use of numerical relativity when dealing with pres- Friedmann universe. With it, we could focus on fluc- sured fields of matter. We indeed also exhibited scenarios tuations of a single barotropic fluid and investigate the where the linear approach fails to give an acceptable ap- critical collapse with respect to the constant parameter proximation of the solution, even at small amplitudes. p ω = of the equation of state. The critical collapse and, more generally, the spheri- e In GR, the question of the observables is always tricky cal collapse are studies that can be declined in numer- because of the gauge dependence of the tensors compo- ous ways by using numerical relativity, depending on the nents. Even such an important quantity as the energy- fluid(s) of matter, the Dark Energy model, the cosmolog- density contrast is gauge dependent because it consists in ical epoch, the scale of interest, the coupling between the a local-background comparison of variables that do not source terms,. . . This article provides lots of perspectives share the same . To avoid difficulties of in- for future works. We give now some of these in what terpretation, we worked first in the synchronous gauge follows. which was good enough for our preliminary observations We used the barotropic linear equation of state p = ωe, despite its well-known instability. But we used the 1+log where e is the energy-density and p is the pressure be- slicing, which was found more appropriated, to simulate cause this is what is commonly considered in cosmology. black holes formations in the second part of the work. However, other equations of state are often used in the Concerning the universality of the critical collapse, we astrophysical case, such as the one for an ideal fluid (37) saw that, for a fixed profile of the energy-density contrast, and the one for a polytropic fluid (38). Threshold values there is a critical value ω∗ of the equation of state pa- of the parameters contained in these equations and the rameter under which the fluctuation collapses to a black associated critical phenomenon could be examined. This hole and upper which it is diluting to the background. would extend again universality with respect to a larger For sub-critical solutions, that is to say collapsing solu- panel of matter species or, on the contrary, lead to the tions, we tried to prove numerically the universal scaling discovery of a breaking of the universality. law for the mass of the formed black hole M ∝ |ω − ω∗|γ As explained in the Introduction and in section V A, similar to what is found in well known critical phenom- the code was designed to evolve possibly two non inter- ena, with a critical exponent independent on the value acting fluids of matter, with a scalar field and a cosmolog- of the equation of state parameter ω. This has been ical constant. In this work, we only explored the case of shown in the Minkowski and the de Sitter cases. For the a single fluid of matter but any other combination of the Einstein-de Sitter universe, we first saw no scaling law, previous ingredients is possible. For example, the late possibly because we could not go as close as necessary to time evolution of pressured matter minimally (or not) the critical solution to observe it. The other possibility coupled to a scalar field would have interesting cosmol- 18 ogy applications. A second obvious application would be start by squaring the second equation of (97) : the study of structure formation at the epoch of matter- 2 i 2 4 2 radiation equivalence since we can evolve conjointly dust S := S Si = (e + p) W v . (98) and radiation, during the inflation era or in a de Sitter background. A third application would be the study of 1 Recalling that v2 = 1 − , (98) becomes voids since a negative energy-density contrast can easily W 2 be implemented. From the numerical relativity point of view, we built S2 = (e + p)2W 4 − (e + p)2W 2. (99) an code which appeared to have some imperfections. Some improvements could be performed, such as the The third equation of (97) gives implementation of an adaptive mesh refinement or a log- arithmic spatial discretization. But the question of the (e + p)2W 4 = (τ + D + p)2 implementation of the cosmological boundary conditions (e + p)2W 2 = (e + p)(τ + D + p). is more tricky and would probably require a research work in itself. Numerical relativity is a subject currently Reinserting in (99) gives the relation[55] on the rise, thanks to the active field of gravitational waves. It is becoming more and more employed within S2 = (τ + D)2 + (τ + D)(p − e) − pe. (100) the frame of cosmology, especially concerning the PBH formation but not only, and we hope to have enhance a Using now the equation of state p = ωe, we obtain a bit the marriage of the two disciplines. quadratic equation in e :

Acknowledgments The authors warmly thanks Dr. I. − ωe2 + (ω − 1)(τ + D)e + (τ + D)2 − S2 = 0. (101) Cordero-Carri´onfor helpful advices and discussions con- cerning the Valencia formulation and the numerical Its solutions are methods used in this work. F. S. is supported by a FRS-FNRS (Belgian Funds for Scientific Research) Re- S2 e = (τ + D) − , if ω = 0, (102) search Fellowship. J.R. is supported by the European τ + D Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (Ad- and vanced Grant agreement No. 670874). This research used resources of the ”Plateforme Technologique de Cal- e = τ + D, if ω = −1, (103) cul Intensif (PTCI)” (http://www.ptci.unamur.be) lo- cated at the University of Namur, Belgium, which is and supported by the F.R.S.-FNRS under the convention p No. 2.5020.11. The PTCI is member of the ”Con- (ω − 1)(τ + D) ± (ω + 1)(τ + D)2 − 4ωS2 e = , sortium des Equipements´ de Calcul Intensif (CECI)”´ ± 2ω (http://www.ceci-hpc.be). if ω 6= 0 (104)

The sign we have to consider depends on the value of ω. Appendix A : Recovering the primitive variables If 0 < ω ≤ 1, we have (ω − 1)(τ + D) ≤ 0 and thus from the conserved ones we have to take the plus to keep a non negative energy density. The conserved variables are defined in such a way : If ω > 1, the positivity of the interior of the root and the fact that  D = ρW 2 2  2 2  2 (ω +1)(τ +D) −4ωS ≤ (ω +1) (τ + D) − S (105) Si = (e + p)W vi (97) τ = (e + p)W 2 − p − D imply that where the rest-mass density ρ, the energy density e, (τ + D)2 − S2 e+e− = ≤ 0, (106) the pressure p, the velocity vi and the Lorentz factor −ω 1 W = √ are the primitive variables. In general, 1 − v2 where the equality holds if and only if e = 0. The latter the inversion of this relation is not analytical and re- case is trivial,because it requires D = 0, Si = 0 and τ = quires a numerical root-finding procedure (see [37] for 0, and in the other cases, this means that the solutions e+ more details). and e− have opposite signs and that only one is positive. In the case of the barotropic equation of state p = ωe, If ω < 0, the situation is less obvious because both this inversion can be made analytically. To obtain it, we solutions can be positive. For example, if ω ∈] − 1; 0[, we 19 have Finally, the auxiliary variable ∆ˆ r is evolved partially im- plicitly : (τ + D)2 − S2 e+e− = −ω 2 4α 2 2 2 4 2 L2 ∆ˆ r = − (Aa∂rα + α∂rAa) − (Aa − Ab) (e + p)W − p − (e + p) W v ( ) aˆ rˆb = " −ω ξα 2 ∂rψ 2 4 2 2 2 + ∂rAa − ∂rK + 6Aa (ω + 1) W + ω − 2ω(ω + 1)W e aˆ 3 ψ = −ω !# 2 ∂ ˆb (ω + 1)2W 4v2e2 + (A − A ) + r , (115) − a b r ˆ −ω b  2 4 2 2 2 2 (ω + 1) W (1 − v ) + ω − 2ω(ω + 1)W e ˆ r ξα = L3(∆ˆ r ) = 2αAa∆ − 8πjr . (116) −ω aˆ > 0, Note that general expressions can be found in [31]. because all the terms of the numerator are positive. This means that both solutions have the same sign and thus are positive. The choice must be done thanks to the con- Appendix C : Kodama mass and mean tinuity of the solution with time, which can be difficult energy-density contrast in BSSN variables numerically. Ones the energy density e is computed, the other vari- The Kodama mass was first defined in [48] but we take ables follow easily : [13] and [47] as references. Recall that, in spherical symmetry, the areal radius is p = ωe (107) the positive quantity R(t, r) defined by the area A(t, r) s τ + D + p of the surface defined by constant t and r coordinates in W = (108) e + p such a way : S 2 v = i (109) A = 4πR . (117) i (e + p)W 2 D In our BSSN metric (2), the areal radius is simply the ρ = . (110) W square root of its θθ component : p √ 2 ˆ R = gθθ = ψ a br. (118) Appendix B : Splitting for the source terms in the PIRK operators   gtt gtr Consider now the 2-metric GAB = with grt grr The evolution equations are written in the form (40) A, B ∈ t, r. We define the Kodama vector by because we are using the PIRK algorithm. The splitting A AB has been chosen to ensure the scheme to be as stable as K =  ∂BR, (119) possible (see [24]). In the first step, the hydrodynamical √ conserved variables, the cosmological scale factor a, the where AB = −GεAB with εAB being the Levi-Civita AB AC BD lapse α, the elements of the conformal 3-metrica ˆ and symbol and  = G G CD. Working in the zero ˆb and ψ are evolved explicitly. These are thus included shift gauge gives in the L1 operator. In the second step, the extrinsic curvature is evolved. This means that K and A are ∂rR a Kt = − √ (120) split into the following L2 and L3 operators [56] : αψ2a aˆ  1  r ∂tR L = − ∇r∇ α − ∇2α K = √ . (121) 2(Aa) r 3 αψ2a aˆ   r 1 A µ +α Rr − R , (111) The tensor K is extended to a 4-vector K by posing 3 θ φ µ µ ν K = K = 0. The quantity S = Tν K is thus a 16π L = αKA − (S − S ) , (112) conserved current (see [48] and [47] for explanations) and 3(Aa) a 3 a b its integral, the Kodama mass, is a conserved quantity. 2 L2(K) = −∇ α, (113) The Kodama mass within a sphere of radius r at time t is thus defined by  1  L = α A2 + 2A2 + K2 3(K) a b 3 Z r t 2 MK (t, r) := 4π S α(t, x)R (t, x)dx. (122) +4πα (E + Sa + 2Sb) . (114) 0 20

By developing St, we find comoving gauge it gives the relation

r p Z 2 ˆ M (t, r) = 4π T t −R2∂ R + T t R2∂ R dx. MK (t, r) − MK (t, ψ br) = K t r r t s 0 Z r ˆ ˆ! (123) b ∂rψ x∂rb 4πa3e δψ6 x2 1 + 2x + dx, (130) The expression (9) gives, in the case of a universe filled 0 aˆ ψ 2ˆb with one fluid of matter (other cases do not change much things), e − e only in term of the energy-density contrast δ = . e t 2 Tt = −(e + p)W + p = −E (124) Concerning the mean energy-density contrast, recall 2 that it is defined in the following way : t W Sr Tr = (e + p) vr = . (125) α α R R 2 0 δR dR δmean(t, r) = . (131) In terms of BSSN variables, we have R R 2 0 R dR s ˆ ˆ! By using BSSN equations, it gives 2 6 3 2 b ∂rψ r∂rb R ∂rR = ψ a r 1 + 2r + (126) r q ˆ  ˆ  aˆ ψ 2ˆb R 6 3 b 2 ∂r ψ x∂r b 0 δψ a aˆ x 1 + 2x ψ + ˆ dx s δ (t, r) = 2b , ! mean q   ˆ ˆ R r ˆb ∂r ψ x∂rˆb 2 6 3 3 b a˙ ∂tψ ∂tb ψ6a3 x2 1 + 2x + dx R ∂ R = ψ a r + 2 + 0 aˆ ψ ˆ t ˆ 2b aˆ a ψ 2b (132) s R3 ˆ   where the denominator can be replaced by = 6 3 3 b K = −αψ a r + Ab , (127) 3 aˆ 3 q ˆ ψ6a3 b r3 aˆ . We thus see the direct relation between the where we have use (11) and (12) for the last equality. In 3 conclusion, the expression for the Kodama mass in BSSN mean energy-density contrast and the compaction in the variables is comoving gauge by looking at the relation (130). s Z r ˆ " ˆ! 3 6 2 b ∂rψ r∂rb MK (t, r) = 4πa ψ x E 1 + 2r + Appendix D : Derivation of the long-wavelength 0 aˆ ψ 2ˆb solution in the BSSN variables # K  −xS + A dx. (128) We assume that, at fixed time, the universe becomes r 3 b locally flat homogeneous and isotropic in the limit  → 0. 2 We thus assume, still following [50], thatγ ˆij = O( ). As The corresponding quantity for the Friedmann universe described in details in [47], we then obtain used as background is thus ψ = O(0), (133) Z r 3 2 4 3 i MK (t, r) = 4πa ex dx = π(ar) e. (129) v = O(), (134) 0 3 3 vi = O( ), (135) i 4 Note that in the definition of the compaction function Div = O( ), (136) (55) we need to compute it at the same areal radius than W = 1 + O(6), (137) the local Kodama mass, that is to say δ = O(2), (138) s ˆ 2 p ˆ Aij = O( ), (139) 2 ˆ 4 3 6 b 3 MK (t, ψ br) = πa ψ e r 2 3 aˆ hij = O( ), (140) √ 2 Z ψ2 ˆbr χ = O( ), (141) 3 2 = 4πa e x dx κ = O(2), (142) 0 s ! where we used the following notations Z r ˆb ∂ ψ y∂ ˆb = 4πa3e ψ6y2 1 + 2y r + r dy, ˆ e − e 0 aˆ ψ 2b δ := , e where we made the change of variable x = hij :=γ ˆij − γˆ , q ij ψ(t, y)2 ˆb(t, y)y for the last equality. The last expres- χ := α − 1, sion, though less simple, can be useful because of its sim- K − K κ := . ilarity with the first term of (128). For example, in a K 21

Moreover, the conformal factor can be decomposed in (149) and (150), by such a way :  2 2 1 4 Aˆij = pijH + O( ), (153) ψ(t, xi) = Ψ(xi) 1 + ξ(t, xi) , (143) 3ω + 5 aH 4  1 2 h = − p + O(4). (154) where Ψ = O(0) and ξ = O(2). In fact, as also shown in ij (3ω + 5)(3ω + 1) ij aH [50] and [47], all slicings coincide up to O(), which allows to apply the long-wavelength scheme in any slicing. After If we want to use spherical symmetry and the BSSN vari- having done it in the special case where p = ωe, the ables presented in section II, we must use the quantities " evolution equations become 1 4  1  p := pr = − ∂2Ψ − ∂ Ψ a r Ψ4 3Ψ r r r δ˙ + 6ξ˙ + 3Hω (χ + κ) = O(4), (144) # 1 4 δ˙ + 6ξ˙ = O(4), (145) + (∂ Ψ)2 , (155) 1 + ω Ψ2 r ∂ a3(1 + ω)eu  = −a3e [ω∂ δ + (1 + ω)∂ χ] " t i i i 1 2  1  5 p := pθ = ∂2Ψ − ∂ Ψ + O( ), (146) b θ Ψ4 3Ψ r r r 5 2 4 ∆Ψ = −2πΨ a e(δ − 2κ) + O( ), (147) # 2 2 −1 3ω − 1 3(1 + ω) 1 + 3ω H κ˙ = κ − χ − δ − 2 (∂rΨ) . (156) 2 2 2 Ψ 4 + O( ), (148) With these definitions, we can write 4 ∂thij = −2Aˆij + O( ), (149) 2  1 2 "   A = p H + O(4), ˆ ˆ 1 2 1 a 3ω + 5 a aH ∂tAij + 3HAij = 2 4 − DiDjΨ − γˆij∆Ψ a Ψ Ψ 3  2 2 1 4 # Ab = pbH + O( ), 6  1 k  3ω + 5 aH + D ΨD Ψ − γˆ D ΨD Ψ + O(4), (150) Ψ2 i j 3 ij k  2 4 1 4 aˆ = 1 − pa + O( ),  6 ˆi  6 6 (3ω + 5)(3ω + 1) aH Di Ψ Aj + 2HΨ Djκ = 8πΨ (1 + ω)euj 4  1 2 + O(5), (151) ˆb = 1 − p + O(4), (3ω + 5)(3ω + 1) b aH where D and ∆ are the covariant derivative and the where we have pa + 2pb = 0, implying the desired condi- ˆ2 4 Laplacian operators related to the flat metric written in tions Aa + 2Ab = 0 anda ˆb = 1 + O( ). The variable the coordinates {xi}. In spherical coordinates, we have ∆ˆ r can be determined thanks to equation (5). 2 2 2 that this metric is given by γˆij = diag(1, r , r sin θ). We will now determine the expressions for δ, ξ, κ and χ The resulting equations have been solved in [47] for by solving equations (144), (145), (147), (148) and (53). the constant mean curvature (CMC) slicing (for which The combination of the first two ones gives K = K), the comoving slicing, the uniform-density slic- −1 ˙ 4 ing (for which δ = 0) and the geodesic slicing. The pa- H δ + 3(1 + ω)(χ + κ) = O( ). (157) per also gives comparison of the solutions between these The Bona-Masso equation (53) reads, after being ex- four slicings. In our case, we will solve them for a gen- panded around α = 1, eral Bona-Masso slicing (53). This solution has, to our knowledge, never been derived in the literature. H−1χ˙ = 3f(1)κ + O(4). (158) We start with the computation of the variables Aˆ ij With equation (148) and the change of variable s = ln a, and h which, as explained in [47], do not depend on the ij we obtain the differential linear system slicing to O(2). If we use the intermediate variable κ κ " 4 1 2  1  ∂s δ = M(ω) δ + O( ), (159) p := − D D Ψ − γˆ ∆Ψ χ χ ij Ψ4 Ψ i j 3 ij # where the matrix M(ω) is given by 6  1 k  + D ΨD Ψ − γˆ D ΨD Ψ ,(152) 2 i j ij k  3ω − 1 1 + 3ω 3(1 + ω) Ψ 3 − −  2 2 2  M(ω) = −3(1 + ω) 0 −3(1 + ω) . (160) the explicit expressions of Aˆij and hij are given, solving 3f(1) 0 0 22

This matrix admits the following eigenvalues : where the integration constant C can be absorbed into Ψ (see [47] in a similar case). λ∗ = 1 + 3ω, (161) The last equation that remains to solve is (146) to find 3 h i the velocity v . By using the expressions of δ and χ, we λ = − 1 + ω ± p(1 + ω)(1 + ω − 8f(1)) .(162) i ± 4 find The general solution of the system (159) is thus given 2 3 by 2 ωv∗ + (1 + ω)v∗ ui = − · 2 1 ∂iF   3ω + 5 (1 + ω)(v∗ − 2v∗) κ Z a   2 λ∗ λ+ λ− 4 1 da˜ 1 δ= C∗v∗a + C+v+a + C−v−a + O( ), (163) · + C + O(5), (172) χ a 0 H(˜a) aH where v∗, v+ and v− are the eigenvectors associated re- where the integration constant C must be set to zero to spectively to λ∗, λ+ and λ− and C∗, C+ and C− are integration constants. keep only growing modes (see [47]). We deduce the value 3(1 + ω) for vi : For the geodesic slicing, λ = − < 0 and + 4 λ− = 0. In the other Bona-Masso slicings of Table III A 3, 2 3  3 λ+ and λ− are complex conjugates with a negative real 2 ωv∗ + (1 + ω)v∗ 1 5 vi = − · 2 1 ∂iF a + O( ). part. Since we only take pure growing modes, we set 3ω + 5 (1 + ω)(v∗ − 2v∗) aH C+ = C− = 0. We have that the eigenvector of λ∗ is (173) given by To summarize the results in terms of BSSN variables  1  2  in spherical symmetry, the long-wavelength solution is v∗ (1 + 3ω) 2 given by v∗ = −3(1 + ω)(1 + 3ω + 3f(1)) . (164) 3 v∗ 3(1 + 3ω)f(1) 2! We can then deduce the two useful relations 1 v2  1  ψ = Ψ 1 − ∗ F 6(1 + ω) v2 − 2v1 aH v1 ∗ ∗ κ = ∗ δ + O(4), (165) 2 +O(4), v∗ v3 2  1 2 ∗ 4 A = p H + O(4), χ = 2 δ + O( ). (166) a a v∗ 3ω + 5 aH 2  1 2 By inserting (165) in (147), we finally obtain A = p H + O(4), b 3ω + 5 b aH v2  1 2  2 ∗ 4 4 1 4 δ = 2 1 F + O( ), (167) aˆ = 1 − pa + O( ), v∗ − 2v∗ aH (3ω + 5)(3ω + 1) aH 4  1 2 where we have defined ˆb = 1 − p + O(4), (3ω + 5)(3ω + 1) b aH 4 ∆Ψ F := − . (168) v2  1 2 3 Ψ5 ∗ 4 δ = 2 1 F + O( ), v∗ − 2v∗ aH We immediately deduce the values of κ, χ and ξ thanks 1  2! to (165), (166) and (145) : v∗ 1 4 K = K 1 + 2 1 F + O( ), v∗ − 2v∗ aH 1 2 v  1  2 ∗ 4 v3  1  κ = 2 1 F + O( ), (169) ∗ 4 v∗ − 2v∗ aH α = 1 + 2 1 F + O( ), v∗ − 2v∗ aH 3 2 v  1  3 ∗ 4 2 ωv2 + (1 + ω)v3  1  χ = 2 1 F + O( ), (170) ∗ ∗ v∗ − 2v∗ aH vr = − · 2 1 ∂iF a 3ω + 5 (1 + ω)(v∗ − 2v∗) aH 2  2 1 v∗ 1 +O(5). ξ = − 2 1 F + C 6(1 + ω) v∗ − 2v∗ aH +O(4), (171) 23

[1] G. Darmois, Les ´equations de la gravitation einsteinienne (1997), ISSN 1538-4357, URL http://dx.doi.org/10. (Gauthier-Villars, 1927). 1086/303604. [2] A. Lichnerowicz, Probl`emesglobaux en m´ecanique rela- [27] T. W. Baumgarte and S. L. Shapiro, Numerical Relativ- tiviste (Hermann et Cie, 1939). ity: Solving Einstein’s Equations on the Computer (Cam- [3] A. Lichnerowicz, J. Math. pures et appl. 23, 37 (1944). bridge Univ. Press, Cambridge, 2010). [4] A. Lichnerowicz, Bulletin de la Soci´et´eMath´ematiquede [28] M. Shibata, Numerical Relativity (World Scientific, Sin- France 80, 237 (1952). gapore, 2016). [5] Y. Four`es-Bruhat, Journal of Rational Mechanics and [29] S. C. Noble and M. W. Choptuik, Physical Review D 93 Analysis 5, 951 (1956). (2016). [6] P. A. M. Dirac, Proceedings of the Royal Society of Lon- [30] I. Cordero-Carrion and P. Cerda-Duran, don. Series A. Mathematical and Physical Sciences 246, https://arxiv.org/abs/1211.5930 (2016). 333 (1958), ISSN 2053-9169, URL http://dx.doi.org/ [31] P. J. Montero and I. Cordero-Carrion, Physical Review 10.1098/rspa.1958.0142. D 85 (2012). [7] P. A. M. Dirac, Physical Review 114, 924 (1959). [32] M. Alcubierre and M. D. Mendez, General Relativity [8] R. Arnowitt, S. Deser, and C. W. Misner, in Gravitation and Gravitation 43, 2769 (2011), ISSN 1572-9532, URL : an introduction to current research (Chap. 7), edited by http://dx.doi.org/10.1007/s10714-011-1202-x. L. Witten (John WIler and Sons Inc, New York, London, [33] M. Alcubierre, Introduction to 3+1 Numerical Relativity 1962), pp. 227–265. (Oxford University Press (OUP), New York, 2008). [9] T. Nakamura, K. Oohara, and Y. Kojima, Prog. Theor. [34] A. Arbona and C. Bona, Computer Physics Com- Phys. Suppl. 90 (1987). munications 118, 229 (1999), ISSN 0010-4655, URL [10] T. Nakamura, in 8th Nishinomiya-Yukawa Memo- http://www.sciencedirect.com/science/article/ rial Symposium : Relativistic Cosmology, edited by pii/S0010465599001915. T. M. Sasaki ed., Universal Academy Press (1994). [35] M. Alcubierre and J. A. Gonz´alez,Computer Physics [11] M. Shibata and T. Nakamura, Physical Review D 52, Communications 167, 76 (2005), ISSN 0010- 5428 (1995). 4655, URL http://www.sciencedirect.com/science/ [12] T. W. Baumgarte and S. L. Shapiro, Physical Review D article/pii/S0010465505000597. 59 (1998), ISSN 1089-4918, URL http://dx.doi.org/ [36] M. Ruiz, M. Alcubierre, and D. N´u˜nez, Gen- 10.1103/PhysRevD.59.024007. eral Relativity and Gravitation 40, 159 (2008), [13] M. Shibata and M. Sasaki, Physical Review D 60 ISSN 1572-9532, URL http://dx.doi.org/10.1007/ (1999), ISSN 1089-4918, URL http://dx.doi.org/10. s10714-007-0522-3. 1103/PhysRevD.60.084002. [37] L. Rezzolla and O. Zanotti, Relativistic Hydrodynamics [14] M. Shibata and K. Uryu, Progress of Theoretical Physics (Oxford University Press (OUP), Oxford, 2013). 107, 265 (2002), ISSN 1347-4081, URL http://dx.doi. [38] P. J. Montero, T. W. Baumgarte, and E. M¨uller,Physical org/10.1143/PTP.107.265. Review D 89 (2014). [15] F. Pretorius, Physical Review Letters 95 (2005). [39] T. W. Baumgarte, P. J. Montero, I. Cordero-Carrion, [16] B. P. A. et al, Physical Review Letters 116 (2016). and E. M¨uller,Physical Review D 87 (2013). [17] M. W. Choptuik, Physical Review Letters 70, 9 (1993). [40] B. V. Leer, Journal of Computational Physics [18] C. Gundlach and J. M. Mart´ın-Garc´ıa,Living Reviews 23, 276 (1977), ISSN 0021-9991, URL http: in Relativity 10 (2007), ISSN 1433-8351, URL http:// //www.sciencedirect.com/science/article/pii/ dx.doi.org/10.12942/lrr-2007-5. 002199917790095X. [19] C. R. Evans and J. S. Coleman, Physical Review Letters [41] A. Harten, P. D. Lax, and B. v. Leer, SIAM Review 72, 1782 (1994). 25, 35 (1983), https://doi.org/10.1137/1025002, URL [20] D. Maison, Physics Letters B 366, 82 (1996), ISSN 0370- https://doi.org/10.1137/1025002. 2693, URL http://www.sciencedirect.com/science/ [42] B. Einfeldt, SIAM Journal on Numerical Analysis 25, 294 article/pii/0370269395013814. (1988), https://doi.org/10.1137/0725021, URL https:// [21] D. W. Neilsen and M. W. Choptuik, Classical and Quan- doi.org/10.1137/0725021. tum 17, 761 (2000), ISSN 1361-6382, URL http: [43] M. Alcubierre, B. Br¨ugmann,P. Diener, M. Koppitz, //dx.doi.org/10.1088/0264-9381/17/4/303. D. Pollney, E. Seidel, and R. Takahashi, Physical Review [22] J. C. Niemeyer and K. Jedamzik, Physical Review D 59 D 67 (2003), ISSN 1089-4918, URL http://dx.doi.org/ (1999). 10.1103/PhysRevD.67.084023. [23] I. Hawke and J. M. Stewart, Classical and Quantum [44] E. Gourgoulhon, 3+1 Formalism in General Relativity Gravity 19, 3687 (2002), ISSN 0264-9381, URL http: : Bases of Numerical Relativity (Springer, Heidelberg, //dx.doi.org/10.1088/0264-9381/19/14/310. 2012). [24] J. Rekier, I. Cordero-Carrion, and A. Fuzfa, Phys. Rev. [45] C. Bona, J. Mass´o, E. Seidel, and J. Stela, Physical Re- D 91, 024025 (2015), 1409.3476, URL https://arxiv. view Letters 75, 600 (1995). org/pdf/1409.3476.pdf. [46] J. M. Torres, M. Alcubierre, A. Diez-Tejedor, and [25] J. Rekier, A. Fuzfa, and I. Cordero-Carrion, Phys. Rev. D. Nu˜nez,Physical Review D 90 (2014). D 93, 043533 (2016), 1509.08354, URL https://arxiv. [47] T. Harada, C.-M. Yoo, T. Nakama, and Y. Koga, Physi- org/pdf/1509.08354.pdf. cal Review D 91 (2015). [26] F. Banyuls, J. A. Font, J. M. Ibanez, J. M. Marti, [48] H. Kodama, Prog. Theor. Phys. 63, 1217 (1980). and J. A. Miralles, The Astrophysical Journal 476, 221 [49] I. Musco, Physical Review D 100 (2019). 24

[50] D. H. Lyth, K. A. Malik, and M. Sasaki, Journal of Cosmology and Astroparticle Physics 2005, 004 (2005), ISSN 1475-7516, URL http://dx.doi.org/10. 1088/1475-7516/2005/05/004. [51] T. Padmanabhan, Structure Formation In The Universe (Cambridge Univ. Press, Cambridge, 1993). [52] I. Musco, J. C. Miller, and A. G. Polnarev, Classical and 26, 235001 (2009), ISSN 1361- 6382, URL http://dx.doi.org/10.1088/0264-9381/ 26/23/235001. [53] I. Musco and J. C. Miller, Classical and Quantum Grav- ity 30, 145009 (2013), ISSN 1361-6382, URL http: //dx.doi.org/10.1088/0264-9381/30/14/145009. [54] We point out here an apparent little typo in [31] where the terms with indices θ are missing. [55] Note that this relation is correct whatever the equation of state and does not requires spherical symmetry. r ˆ r [56] The terms of Rr and R which are proportional to ∆ and ˆ r ∂r∆ are in fact included in the L3(Aa) operator instead

of L2(Aa).