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

Spherical scalar collapse in a type-II minimally modified gravity

Atabak Fathe Jalali [email protected] KTH Royal Institute of Technology, SE-100 44 Stockholm, Sweden Center for Gravitational Physics and Quantum Information (CGPQI), Yukawa Institute for Theoretical Physics (YITP), Kyoto University, 606-8502, Kyoto, Japan    Paul Martens [email protected] Center for Gravitational Physics and Quantum Information (CGPQI), Yukawa Institute for Theoretical Physics (YITP), Kyoto University, 606-8502, Kyoto, Japan    Shinji Mukohyama [email protected] Center for Gravitational Physics and Quantum Information (CGPQI), Yukawa Institute for Theoretical Physics (YITP), Kyoto University, 606-8502, Kyoto, Japan Kavli Institute for the Physics and Mathematics of the Universe (WPI), The University of Tokyo, Kashiwa, Chiba 277-8583, Japan
(February 12, 2025)
Abstract

We investigate the spherically-symmetric gravitational collapse of a massless scalar field in the framework of a type-II minimally modified gravity theory called VCDM. This theory propagates only two local physical degrees of freedom (DoF) supplemented by the so-called instantaneous (or shadowy) mode. Imposing asymptotically flat spacetime in the standard Minkowski time slicing, one can integrate out the instantaneous mode. Consequently, the equations of motion reduce to those in general relativity (GR) with the maximal slicing. Unlike GR, however, VCDM lacks 4D diffeomorphism invariance, and thus one cannot change the time slicing that is preferred by the theory. We then numerically evolve the system to see if and how a black hole forms. For small amplitudes of the initial scalar profile, we find that its collapse does not generate any black hole, singularity or breakdown of the time slicing. For sufficiently large amplitudes, however, the collapse does indeed result in the formation of an apparent horizon in a finite time. After that, the solution outside the horizon is described by a static configuration, i.e. the Schwarzschild geometry with a finite and time-independent lapse function. Inside the horizon, on the other hand, the numerical results indicate that the lapse function keeps decreasing towards zero so that the central singularity is never reached. This implies the necessity for a UV completion of the theory to describe physics inside the horizon. Still, we can conclude that VCDM is able to fully describe the entire time evolution of the Universe outside the black hole horizon without knowledge about such a UV completion.

preprint: YITP-23-75, IPMU23-0023

I Introduction

When general relativity (GR) was first formulated, it showed promise by solving observational riddles that deprived physicists of their sleep at the time. Even today, GR continues to produce physically accurate predictions, e.g. the detection of gravitational waves (GWs) by LIGO/Virgo [1, 2] and the first-ever photograph of a black hole shadow [3]. However, albeit a simple yet rigorous, elegant, and symmetrically satisfying theory, GR has proven incomplete on quantum gravity scales, and some observations may currently be bearing witness to its failure on cosmological scales. Once more, physicists find themselves in a state of sleep deprivation.

This motivates us to look beyond GR. Do we need a completely new theory? Is it possible to modify gravity in a way such that it includes a complete description of quantum gravity, and breeds a cosmological framework that is consistent with observations? The experiments conducted at LIGO/Virgo open up the possibility of probing strongly gravitating events through GWs, such as black hole binaries and gravitational collapse. This allows for further testing of the validity of GR and alternative frameworks.

Several such alternative theories of gravity have already been proposed. In the high-energy limit, superstring theories [4] and Hořava-Lifshitz gravity [5] are examples of candidates for a theory of quantum gravity, while massive gravity, bimetric gravity, and various scalar-tensor theories [6, 7, 8, 9, 10] carry implications for the dark sector of the Universe as well as its accelerated expansion. A common artifact when modifying gravity is accompanying DoF. On cosmological scales, these are exploited to explain observational discrepancies. However, on astrophysical scales, these must also avoid contradicting well-established experimental data and the appearance of various instabilities such as ghosts [11]. One common method of dealing with this is to implement different screening mechanisms [12]. Alternatively, one could consider modifying gravity minimally, i.e. keeping only two local physical DoF. Such a theory is denoted a minimally modified gravity (MMG) theory [13]. This class of theories defines itself by not introducing any additional local physical degrees of freedom other than those in GR. Therefore, they easily avoid the aforementioned instabilities and constraints that could come from the extra propagating degrees of freedom that are commonly found in other modified gravity theories. And this is the case even without needing any screening mechanisms.

It was previously established in [14] that all MMG theories may be divided into two types: type-I and type-II. The former are theories that are equivalent to GR in the absence of matter but that modify gravity due to non-trivial matter coupling [15], while the latter are theories simply different from GR, even in the absence of matter. For example, the class of MMGs studied in [13] were, by [16], mostly shown to be of type-I, and the most generic construction of a subclass of type-I MMG with the standard dispersion relation for the tensorial GWs is elaborated upon in subsection IV A of [17]. Another example would be [18]. Instances of type-II MMG theories include the Cuscuton [19], the minimal theory of massive gravity [20], a consistent D4D\to 4 Einstein-Gauss-Bonnet gravity [21], and VCDM [22] (see also [23, 24]).

We have chosen to further investigate VCDM [22]. Unlike GR, VCDM does not have full 4D diffeomorphism invariance (although 3D spatial diffeomorphism invariance still holds) and effectively substitutes the cosmological constant with a free potential function V(ϕ)V(\phi) of a non-propagating auxiliary scalar field ϕ\phi in the gravitational action. This last feature gives “VCDM” its name. The theory is kept minimal thanks to the built-in constraints that are imposed on the field ϕ\phi. Consequently, VCDM is ghost-free, as the only propagating modes are the standard tensor modes, without requiring any screening mechanisms are needed. Nonetheless, the theory does admit a phenomenology that is different from GR. Indeed, so far, VCDM has proven to be a fruitful framework within which to approach multiple questions, such as tensions in late-time cosmology [22, 25, 26] and a bouncing Universe [27]. Stars [28], as well as black holes and gravitational collapse [29, 28] have also been studied. A thorough comparison with the Cuscuton has been performed in [17, 30], the latter showing how any solution of Cuscuton is a solution of VCDM, while the converse is not necessarily true. It was further shown that there exists a subset of solutions in VCDM that coincide with exact solutions in GR, given specific conditions on the foliation of spacetime.

In this paper, we are interested in numerically simulating the gravitational collapse of a massless scalar field in VCDM. It is well-established in GR that black holes form after a gravitational collapse. By virtue of its aforementioned overlap with GR solutions, one may expect VCDM to admit black holes from gravitational collapse as well, at least in specific foliations of spacetime. On the other hand, unlike GR, the fundamental symmetry of VCDM is not the 4D diffeomorphism invariance but the invariance under the foliation-preserving diffeomorphism, and thus one cannot change the time slicing that is preferred by the theory. In VCDM, different time foliations represent physically different solutions, and whether or not an apparent horizon (AH) appears before a singularity and a breakdown of the time foliation carries great importance. A collapsing cloud of dust was recently subject to a numerical study and did indeed result in the formation of a black hole [31], its solution coinciding with a particular foliation of the Oppenheimer-Snyder collapse in which the AH forms prior to the singularity and the breakdown of the foliation.

The structure of this paper is as follows. Section II gives a brief overview of VCDM as a theory of gravity and formulate the total action in the unitary gauge. In section III, assuming a spherically symmetric ansatz, we derive the equations of motion. These are simplified by demanding an asymptotically flat spacetime, allowing us to integrate out the instantaneous (or shadowy) mode and acquire a constant ϕ\phi and traceless extrinsic curvature. After specifying relevant boundary conditions and initial conditions, the equations of motion are ready to be integrated. The numerical setup, as well as the simulation results, are presented in section IV. We there confirm the correct behavior of the numerical convergence, the appearance of an AH, and the evolution of the lapse function inside the AH towards zero. Finally, section V provides a summary of the paper and a discussion of the obtained results, concluding with some suggestions for future projects.

II An overview of VCDM

We consider a type-II MMG theory called VCDM, which was first formulated in [22]111In particular, section 2 for the complete construction of VCDM, and section 5 for its phenomenology that differs from GR. The construction of VCDM begins with a canonical transformation of vacuum GR in the ADM formalism. The generating functional FF for this canonical transformation is expressed in terms of two quantities ϕ\phi and ψ\psi. These are equal to some combinations of the transformed and untransformed canonical coordinates. They are later promoted to non-dynamical fields and constrained to evaluate to their respective expressions of the canonical coordinates. A cosmological constant term is then introduced in the frame after the transformation. This demotes the Hamiltonian constraint from first-class to second-class and the theory is now different from GR. However, as a consequence of the demotion, the number of the phase space DoF increases to 55, and the addition of a gauge-fixing term to the action is necessary so as to reduce the dimension of the constraint surface in phase space down to 44 again. The broken symmetry that was originally generated by the Hamiltonian constraint is time diffeomorphism invariance; in exchange, the theory is different from GR, yet propagates only two physical DoF. The theory is also ghost-free, simply because there is no extra gravitational mode besides the standard tensor modes. Thus, it avoids having to employ screening mechanisms, and yet we still have new phenomenology on cosmological scales.

In the ADM formalism, spacetime is foliated by a family of tt-constant non-intersecting 3D space-like hypersurfaces. Choosing a set of coordinates (t,xi)(t,x^{i}) (i=1,2,3i=1,2,3), the line element takes the form

gμνdxμdxν=N2dt2+γij(Nidt+dxi)(Njdt+dxj),g_{\mu\nu}dx^{\mu}dx^{\nu}=-N^{2}dt^{2}+\gamma_{ij}(N^{i}dt+dx^{i})(N^{j}dt+dx^{j}), (1)

where NN is the lapse function, NiN^{i} is the shift vector, and γij\gamma_{ij} (with inverse γij\gamma^{ij}) is the induced 3D spatial metric on the constant tt hypersurface. In contrast to GR and as a consequence of the theory’s construction, the Lagrangian density of VCDM additionally depends on an auxiliary scalar field ϕ\phi, a free function V(ϕ)V(\phi), and two Lagrange multipliers λ\lambda and λgfi\lambda_{\rm gf}^{i}. By working in the so-called unitary gauge, the gravitational action of VCDM can explicitly be written

Ig=MPl2𝑑td3xNγ{12[R+KijKijK22V(ϕ)]λgfiNiϕ34λ2λ(K+ϕ)},I_{\rm g}=M_{\rm Pl}^{2}\int dtd^{3}x\,N\sqrt{\gamma}\left\{\frac{1}{2}\,\left[R+K_{ij}K^{ij}-K^{2}-2V(\phi)\right]-\frac{\lambda_{{\rm gf}}^{i}}{N}\,\partial_{i}\phi-\frac{3}{4}\,\lambda^{2}-\lambda\,(K+\phi)\right\}\,, (2)

where MPl=1/8πGNM_{\text{Pl}}=1/\sqrt{8\pi G_{N}} is the reduced Planck mass, GNG_{N} denotes Newton’s gravitational constant, γ\gamma is the determinant of γij\gamma_{ij}, and RR denotes the three-dimensional Ricci scalar of γij\gamma_{ij}. The extrinsic curvature KijK_{ij} of the tt-constant hypersurfaces, its inverse KijK^{ij}, and its trace KK are defined by

Kij\displaystyle K_{ij} =12N(tγijDiNjDjNi),\displaystyle=\dfrac{1}{2N}(\partial_{t}\gamma_{ij}-D_{i}N_{j}-D_{j}N_{i}), (3)
Kij\displaystyle K^{ij} =γikγjlKkl,\displaystyle=\gamma^{ik}\gamma^{jl}K_{kl}, (4)
K\displaystyle K =γijKij.\displaystyle=\gamma^{ij}K_{ij}. (5)

Here, DiD_{i} is the three-dimensional covariant derivative compatible with γij\gamma_{ij}, and Ni=γijNjN_{i}=\gamma_{ij}N^{j}. Finally, we mention that the two Lagrange multipliers λ\lambda and λgfi\lambda^{i}_{\rm gf} in equation (2) serve to constrain the field ϕ\phi so that the theory remains minimal, i.e. having only two local physical DoF.

As a matter source, we consider a canonical massless scalar field ψ\psi described by the action

Im=MPl22𝑑td3xNγ[(ψ)2γijiψjψ],1N(tNii).I_{\rm m}=\frac{M_{\rm Pl}^{2}}{2}\int dtd^{3}xN\sqrt{\gamma}\left[(\partial_{\perp}\psi)^{2}-\gamma^{ij}\partial_{i}\psi\partial_{j}\psi\right]\,,\quad\partial_{\perp}\equiv\frac{1}{N}(\partial_{t}-N^{i}\partial_{i})\,. (6)

Combining equations (2) and (6) produces the total action

Itot=Ig+Im.I_{\rm tot}=I_{\rm g}+I_{\rm m}\,. (7)

III Setup

This section is dedicated to deriving the equations of motion and preparing these for the numerical integration (section IV). We begin in subsection III.1 by assuming a spherically symmetric ansatz and then deriving the equations of motion. In subsection III.2, we impose asymptotic flatness in the standard Minkowski slicing and integrate out the so-called instantaneous (or shadowy) mode. Consequently, all solutions of this system will be GR solutions in the maximal slicing (i.e. K=0K=0). Subsection III.3 downgrades QQ to a non-dynamical variable, which is practical for numerical stability. We finally conclude by specifying the boundary conditions and initial conditions for all variables in subsections III.4 and III.5, respectively.

III.1 Basic equations

We adopt the spherically symmetric ansatz

N=α(t,r),Nidxi=β(t,r)dr,γijdxidxj=dr2+Φ(t,r)2dΩ22,ψ=ψ(t,r).N=\alpha(t,r)\,,\quad N_{i}dx^{i}=\beta(t,r)dr\,,\quad\gamma_{ij}dx^{i}dx^{j}=dr^{2}+\Phi(t,r)^{2}d\Omega_{2}^{2}\,,\quad\psi=\psi(t,r)\,. (8)

where α\alpha and β\beta now relate to the lapse and the shift, respectively, and Φ\Phi is the areal radius. In parallel, we also write

ϕ=ϕ(t,r),λ=λ(t,r),λgfii=λ~(t,r)r,\phi=\phi(t,r)\,,\quad\lambda=\lambda(t,r)\,,\quad\lambda_{{\rm gf}}^{i}\partial_{i}=\tilde{\lambda}(t,r)\partial_{r}\,, (9)

where dΩ22d\Omega_{2}^{2} is the metric of the unit 22-sphere. In order to simplify the equations of motion, we will use the trace of the extrinsic curvature that now reads

K=2lnΦ1αrβ,K=2\partial_{\perp}\ln\Phi-\frac{1}{\alpha}\partial_{r}\beta\,, (10)

where =(1/α)(tβr)\partial_{\perp}=(1/\alpha)(\partial_{t}-\beta\partial_{r}). Writing KjiγikKkjK^{i}_{j}\equiv\gamma^{ik}K_{kj}, we also introduce the following variables

QKrr13K=23(lnΦ+1αrβ),Pψ,arlnα.Q\equiv K^{r}_{r}-\frac{1}{3}K=-\frac{2}{3}\left(\partial_{\perp}\ln\Phi+\frac{1}{\alpha}\partial_{r}\beta\right)\,,\quad P\equiv\partial_{\perp}\psi\,,\quad a\equiv\partial_{r}\ln\alpha\,. (11)

Given this ansatz, the position of an AH r=rAHr=r_{\rm AH}, can be found by solving222A derivation of this condition can be found in appendix A. For more details, please consult subsection III B in [31] and/or subchapters 2.4 and 5.1.7 in [32].

gΦΦ|r=rAH=0,\left.g^{\Phi\Phi}\right|_{r=r_{\rm AH}}=0\,, (12)

where

gΦΦ(Φ)2+(rΦ)2=(12Q+13K)2Φ2+(rΦ)2.g^{\Phi\Phi}\equiv-(\partial_{\perp}\Phi)^{2}+(\partial_{r}\Phi)^{2}=-\left(\frac{1}{2}Q+\frac{1}{3}K\right)^{2}\Phi^{2}+(\partial_{r}\Phi)^{2}\,. (13)

In terms of the introduced variables, the equations of motion can then be divided into two sets: one set with the dependence on the Lagrange multipliers λ\lambda and λ~\tilde{\lambda} removed, and the other set determining λ\lambda and λ~\tilde{\lambda}.

The first set of equations of motion, i.e. those independent of the Lagrange multipliers λ\lambda and λ~\tilde{\lambda}, consists of constraints, dynamical equations, and non-dynamical equations. The constraints are

r2ΦΦ\displaystyle\frac{\partial_{r}^{2}\Phi}{\Phi} =1(rΦ)22Φ238Q214P214(rψ)212V(ϕ)+16ϕ2,\displaystyle=\frac{1-(\partial_{r}\Phi)^{2}}{2\Phi^{2}}-\frac{3}{8}Q^{2}-\frac{1}{4}P^{2}-\frac{1}{4}(\partial_{r}\psi)^{2}-\frac{1}{2}V(\phi)+\frac{1}{6}\phi^{2}\,, (14)
rQ\displaystyle\partial_{r}Q =3QrlnΦ+Prψ,\displaystyle=-3Q\partial_{r}\ln\Phi+P\partial_{r}\psi\,, (15)
rϕ\displaystyle\partial_{r}\phi =0,\displaystyle=0\,, (16)

where equation (14) comes from the Hamiltonian constraint, equation (15) from the momentum constraint, and equation (16) follows from the variation with respect to λ~\tilde{\lambda}. The dynamical equations are

tψ\displaystyle\partial_{t}\psi =αP+βrψ,\displaystyle=\alpha P+\beta\partial_{r}\psi\,, (17)
tP\displaystyle\partial_{t}P =α[KP+r2ψ+(a+2rlnΦ)rψ]+βrP,\displaystyle=\alpha\left[-KP+\partial_{r}^{2}\psi+(a+2\partial_{r}\ln\Phi)\partial_{r}\psi\right]+\beta\partial_{r}P\,, (18)
tΦ\displaystyle\partial_{t}\Phi =α(13K12Q)Φ+βrΦ,\displaystyle=\alpha\left(\frac{1}{3}K-\frac{1}{2}Q\right)\Phi+\beta\partial_{r}\Phi\,, (19)
tQ=α[KQ14Q2+23(a2+raarlnΦ)+1(rΦ)2Φ216P2+12(rψ)2+19ϕ213V(ϕ)]+β(Prψ3QrlnΦ),\displaystyle\begin{split}\partial_{t}Q&=\alpha\left[-KQ-\frac{1}{4}Q^{2}+\frac{2}{3}(a^{2}+\partial_{r}a-a\partial_{r}\ln\Phi)+\frac{1-(\partial_{r}\Phi)^{2}}{\Phi^{2}}-\frac{1}{6}P^{2}+\frac{1}{2}(\partial_{r}\psi)^{2}\right.\\ &\quad\qquad\left.+\frac{1}{9}\phi^{2}-\frac{1}{3}V(\phi)\right]+\beta(P\partial_{r}\psi-3Q\partial_{r}\ln\Phi),\end{split} (20)
tϕ\displaystyle\partial_{t}\phi =α[a2ra2arlnΦ+32Q2+P2+13ϕ2V(ϕ)].\displaystyle=\alpha\left[-a^{2}-\partial_{r}a-2a\partial_{r}\ln\Phi+\frac{3}{2}Q^{2}+P^{2}+\frac{1}{3}\phi^{2}-V(\phi)\right]\,. (21)

Two non-dynamical equations are derived from the definitions of (a,Q,K)(a,Q,K) and give the first spatial derivative of the lapse and the shift as

rlnα\displaystyle\partial_{r}\ln\alpha =a,\displaystyle=a\,, (22)
rβ\displaystyle\partial_{r}\beta =α(Q+13K).\displaystyle=-\alpha\left(Q+\frac{1}{3}K\right)\,. (23)

The remaining non-dynamical equations are obtained by requiring that the time derivatives of the constraints be consistent with the spatial derivatives of the dynamical equations as

r2K\displaystyle\partial_{r}^{2}K =2(a+rlnΦ)rK[a2+ra+2arlnΦ32Q2P213ϕ2+V(ϕ)][K+ϕ32V(ϕ)],\displaystyle=-2(a+\partial_{r}\ln\Phi)\partial_{r}K-\left[a^{2}+\partial_{r}a+2a\partial_{r}\ln\Phi-\frac{3}{2}Q^{2}-P^{2}-\frac{1}{3}\phi^{2}+V(\phi)\right]\left[K+\phi-\frac{3}{2}V^{\prime}(\phi)\right]\,, (24)
r2a=3ara+a[(rΦ)21Φ2+94Q2+32P2+12(rψ)2]2a2rlnΦa3+P(3Qrψ+2rP)9Q2rlnΦ+2(arlnΦra)rlnΦ.\displaystyle\begin{split}\partial_{r}^{2}a&=-3a\partial_{r}a+a\left[\frac{(\partial_{r}\Phi)^{2}-1}{\Phi^{2}}+\frac{9}{4}Q^{2}+\frac{3}{2}P^{2}+\frac{1}{2}(\partial_{r}\psi)^{2}\right]-2a^{2}\partial_{r}\ln\Phi-a^{3}\\ &\qquad+P(3Q\partial_{r}\psi+2\partial_{r}P)-9Q^{2}\partial_{r}\ln\Phi+2(a\partial_{r}\ln\Phi-\partial_{r}a)\partial_{r}\ln\Phi\,.\end{split} (25)

Finally, the extra equations that determine the Lagrange multipliers λ\lambda and λ~\tilde{\lambda} are

λ=23(K+ϕ),rλ~+2λ~rlnΦ=[23(K+ϕ)V(ϕ)]α.\lambda=-\frac{2}{3}(K+\phi)\,,\quad\partial_{r}\tilde{\lambda}+2\tilde{\lambda}\partial_{r}\ln\Phi=-\left[\frac{2}{3}(K+\phi)-V^{\prime}(\phi)\right]\alpha\,. (26)

We are interested in evolving the nine variables (ψ,P,Φ,Q,ϕ,α,β,K,a)(\psi,P,\Phi,Q,\phi,\alpha,\beta,K,a), and note that λ\lambda and λ~\tilde{\lambda} do not appear in any of the equations (14)-(25). Hence, determining them will not be necessary for the later numerical integration; we shall not consider them anymore.

III.2 Integrating out the shadowy mode

In actual numerical studies, we restrict our considerations to the situation where the solution far from the center approaches a flat spacetime in the standard Minkowski slicing without excitations of the scalar field ψ\psi. In vacuum (or for ψ=const.\psi=\text{const.}), the theory admits a flat spacetime in the standard Minkowski slicing without spontaneous breaking of the spatial diffeomorphism invariance (and thus with λ~=0\tilde{\lambda}=0) if and only if ϕ=ϕ0\phi=\phi_{0}, where ϕ0\phi_{0} is a constant that simultaneously satisfies

V(ϕ0)\displaystyle V^{\prime}(\phi_{0}) =23ϕ0,\displaystyle=\frac{2}{3}\phi_{0}\,, (27)
V(ϕ0)\displaystyle V(\phi_{0}) =13ϕ02.\displaystyle=\frac{1}{3}\phi_{0}^{2}\,. (28)

One can consider equation (27) as the definition of ϕ0\phi_{0} and equation (28) as the condition setting the effective cosmological constant to zero. The latter can be achieved by tuning a constant in V(ϕ)V(\phi). Since equation (16) says that ϕ\phi is independent of rr, the assumed asymptotic condition that the solution should approach the flat spacetime far from the center implies that ϕ=ϕ0\phi=\phi_{0} everywhere, where ϕ0\phi_{0} is defined by equation (27), and that the effective cosmological constant should be set to zero as in equation (28).

Upon introducing this change into equations (14)-(25), the equations of motion simplify. We note that the previously dynamical equation (21) for ϕ\phi downgrades to a non-dynamical one for aa, explicitly

ra=a22arlnΦ+32Q2+P2.\partial_{r}a=-a^{2}-2a\partial_{r}\ln\Phi+\frac{3}{2}Q^{2}+P^{2}\,. (29)

Because of this, we no longer include equation (25) in the set of equations to be solved; it automatically follows from other equations.

Furthermore, the equation (24) for KK simplifies to

r2K=2(a+rlnΦ)rK,\partial_{r}^{2}K=-2(a+\partial_{r}\ln\Phi)\partial_{r}K\,, (30)

which can be integrated once to give rK(t,r)=f1(t)/[α(t,r)Φ(t,r)]\partial_{r}K(t,r)=f_{1}(t)/[\alpha(t,r)\Phi(t,r)], where f1(t)f_{1}(t) is an arbitrary function of tt. The regularity of the solution at the center of spherical symmetry, where Φ(t,r)=0\Phi(t,r)=0, requires that α(t,r=0)0\alpha(t,r=0)\neq 0 and that rK(t,r=0)=0\partial_{r}K(t,r=0)=0. Therefore, the regularity of the center sets f1(t)=0f_{1}(t)=0 and thus rK(t,r)=0\partial_{r}K(t,r)=0 and K(t,r)=f0(t)K(t,r)=f_{0}(t), where f0(t)f_{0}(t) is an arbitrary function of tt. As stated previously, we assume that the solution far from the center approaches a flat spacetime in the standard Minkowski slicing without excitations of the scalar field ψ\psi. This in particular implies that K(t,r)K(t,r) should vanish at the spatial infinity, meaning that f0(t)=0f_{0}(t)=0 and that K(t,r)=0K(t,r)=0. In this way we have successfully integrated out the so-called shadowy mode, avoiding the corresponding boundary value problem.

We also point out that, as KK has now reduced to a constant (actually zero), equation (13) can be re-written

gΦΦ(Φ)2+(rΦ)2=14Q2Φ2+(rΦ)2.g^{\Phi\Phi}\equiv-(\partial_{\perp}\Phi)^{2}+(\partial_{r}\Phi)^{2}=-\frac{1}{4}Q^{2}\Phi^{2}+(\partial_{r}\Phi)^{2}\,. (31)

In addition, KK vanishing implies from equation (26) that λ\lambda also becomes 0. Accordingly, the solutions to the equations (14)-(25) after this simplification will, in fact, be exact GR solutions; we are interested in whether or not an AH appears, and if it does, before or after a breakdown of the time foliation and formation of a singularity. We reiterate that in GR, a change of the spacetime foliation could reverse the order of these events. However, different foliations in VCDM correspond to physically different solutions. Any singularity or breakdown of the time foliation appearing in this theory cannot be circumvented by a change of the time slicing.

III.3 Nondynamical QQ

For actual numerical simulation, one can downgrade QQ from a dynamical variable to a non-dynamical one, abandoning equation (20) and instead using equation (15) to determine QQ. The simplifications in the previous subsection along with such a downgrade give the equations of motion in their final form. Only the following constraint equation remains

r2ΦΦ=1(rΦ)22Φ238Q214P214(rψ)2,\frac{\partial_{r}^{2}\Phi}{\Phi}=\frac{1-(\partial_{r}\Phi)^{2}}{2\Phi^{2}}-\frac{3}{8}Q^{2}-\frac{1}{4}P^{2}-\frac{1}{4}(\partial_{r}\psi)^{2}\,,\\ (32)

while the dynamical equations reduce to

tψ\displaystyle\partial_{t}\psi =αP+βrψ,\displaystyle=\alpha P+\beta\partial_{r}\psi\,, (33)
tP\displaystyle\partial_{t}P =α[r2ψ+(a+2rlnΦ)rψ]+βrP,\displaystyle=\alpha\left[\partial_{r}^{2}\psi+(a+2\partial_{r}\ln\Phi)\partial_{r}\psi\right]+\beta\partial_{r}P\,, (34)
tΦ\displaystyle\partial_{t}\Phi =12αQΦ+βrΦ,\displaystyle=-\frac{1}{2}\alpha Q\Phi+\beta\partial_{r}\Phi\,, (35)

and the non-dynamical equations become

rlnα\displaystyle\partial_{r}\ln\alpha =a,\displaystyle=a\,, (36)
rβ\displaystyle\partial_{r}\beta =αQ,\displaystyle=-\alpha Q\,, (37)
ra\displaystyle\partial_{r}a =a22arlnΦ+32Q2+P2,\displaystyle=-a^{2}-2a\partial_{r}\ln\Phi+\frac{3}{2}Q^{2}+P^{2}\,, (38)
rQ\displaystyle\partial_{r}Q =3QrlnΦ+Prψ,\displaystyle=-3Q\partial_{r}\ln\Phi+P\partial_{r}\psi\,, (39)

where equation (39) comes from equation (15). For convenience, equation (32) can be reformulated to define the quantity 𝒞\mathcal{C} as

𝒞r2Φ+(rΦ)212Φ+[38Q2+14P2+14(rψ)2]Φ.\mathcal{C}\equiv\partial_{r}^{2}\Phi+\frac{(\partial_{r}\Phi)^{2}-1}{2\Phi}+\left[\frac{3}{8}Q^{2}+\frac{1}{4}P^{2}+\frac{1}{4}(\partial_{r}\psi)^{2}\right]\Phi\,. (40)

In this case, if equation (32) is indeed well respected, the quantity 𝒞\mathcal{C} remains identically null. Any deviations from 0, which is bound to happen once numerically implemented, will allow us to watch and quantify the convergence of the numerical integration (see section IV).

III.4 Boundary conditions

We need to be careful about the boundary condition at the center of spherical symmetry, where Φ=0\Phi=0. By redefinition of the radial coordinate rr, we set

Φ(t,r=0)=0,rΦ(t,r=0)>0\Phi(t,r=0)=0\,,\quad\partial_{r}\Phi(t,r=0)>0 (41)

so that the center of spherical symmetry is at r=0r=0 and that we consider the region r0r\geq 0. For the regularity of the center, we require (ψ,P,Q,α\psi,P,Q,\alpha) to be even functions of rr and (Φ,β,a)(\Phi,\beta,a) to be odd functions of rr. Moreover, equations (32) and (39) respectively imply that

rΦ(t,r=0)=1,Q(t,r=0)=0.\partial_{r}\Phi(t,r=0)=1\,,\qquad Q(t,r=0)=0\,. (42)

By Taylor expanding the variables (ψ,P,Φ,Q,α,β,a)(\psi,P,\Phi,Q,\alpha,\beta,a) with respect to rr around r=0r=0 and inserting these, one can confirm that the right-hand sides of all relevant equations (32)-(39) are well-defined in the limit r+0r\to+0, implying that they can also be Taylor expanded with respect to rr. As a consistency check, one can show that the right-hand sides of equations (32), (33), (34), (37), and (38) are even functions of rr, while the right-hand sides of equations (35), (36), and (39) are odd functions of rr. Furthermore, by using the r+0r\to+0 limits of equations (37) and (32) respectively, one can also show that the first rr-derivative of the right-hand side of equation (35) vanishes at r=0r=0.

We now introduce an outer boundary at r=rbr=r_{\rm b} and hereafter consider the region 0rrb0\leq r\leq r_{\rm b}. In order to evolve (ψ,P,Φ)(\psi,P,\Phi) with the dynamical equations (33)-(35), we need to impose appropriate boundary conditions on the outer boundary as well. For example, if rbr_{\rm b} is large enough then we can set

ψ(t,rb)=0,P(t,rb)=0,r2Φ(t,rb)=0.\psi(t,r_{\rm b})=0\,,\quad P(t,r_{\rm b})=0\,,\quad\partial_{r}^{2}\Phi(t,r_{\rm b})=0\,. (43)

Up to here, the overall normalization of α\alpha is not fixed, corresponding to the fact that the theory enjoys time reparametrization symmetry. We fix it by demanding that

α(t,r=rb)=1\alpha(t,r=r_{\rm b})=1 (44)

so that the time tt agrees with the proper time measured by an observer at the outer boundary. The equations (37) and (38) for β\beta and the acceleration aa do not require additional boundary conditions; the fact that β\beta and aa are odd functions of rr already imposes β(t,r=0)=0=a(t,r=0)\beta(t,r=0)=0=a(t,r=0).

III.5 Initial conditions

We consider the following initial conditions at the initial time t=t0t=t_{0}:

P(t0,r)=0,Q(t0,r)=0,α(t0,r)=1,β(t0,r)=0,a(t0,r)=0,P(t_{0},r)=0\,,\quad Q(t_{0},r)=0\,,\quad\alpha(t_{0},r)=1\,,\quad\beta(t_{0},r)=0\,,\quad a(t_{0},r)=0\,, (45)

and

ψ(t0,r)=ψ0(r),Φ(t0,r)=Φ0(r),\psi(t_{0},r)=\psi_{0}(r)\,,\quad\Phi(t_{0},r)=\Phi_{0}(r)\,, (46)

where ψ0(r)\psi_{0}(r) is a free function of rr, which we, for concreteness, take to be an even Gaussian wave packet of the form

ψ0(r)=Aexp[(r2r02)2s4].\psi_{0}(r)=A\exp\left[-\frac{(r^{2}-r_{0}^{2})^{2}}{s^{4}}\right]\,. (47)

The constants AA, r0r_{0} and ss determine the amplitude, starting position, and width of the initial wave packet respectively. Having defined ψ0\psi_{0}, the constraint (32) implies that Φ0(r)\Phi_{0}(r) is the solution to

r2Φ0Φ0=1(rΦ0)22Φ0214(rψ0)2\frac{\partial_{r}^{2}\Phi_{0}}{\Phi_{0}}=\frac{1-(\partial_{r}\Phi_{0})^{2}}{2\Phi_{0}^{2}}-\frac{1}{4}(\partial_{r}\psi_{0})^{2} (48)

with the boundary conditions given in subsection III.4, i.e.

Φ0(r=0)=0,rΦ0(r=0)=1.\Phi_{0}(r=0)=0\,,\quad\partial_{r}\Phi_{0}(r=0)=1\,. (49)

IV Numerical integration

This section details the numerical approach to the problem. Subsection IV.1 begins by explaining the schematics of the integration, explicitly stating the parameter choice for the initial conditions in subsection III.5, as well as different numerical remedies for taming arising instabilities. In subsection IV.2 we present the simulation results and confirm, in subsection IV.3, the appearance of an apparent horizon followed by the evolution of the lapse function inside the horizon towards zero. Finally, we discuss the implications of the initial condition parameter choice on the dynamics and end results of the collapse in subsection IV.4.

IV.1 Numerical setup

The dynamics of the collapse are fully described by three dynamical equations for (ψ,P,Φ)(\psi,P,\Phi) and four non-dynamical ones for (α,β,a,Q)(\alpha,\beta,a,Q)333These form a strongly hyperbolic system of partial differential equations, since after having integrated out the shadowy mode, they were shown to be the same as in GR.. We solve the system described in subsection III.3 with the boundary conditions and initial conditions formulated in subsections III.4 and III.5. In order to remain consistent, the non-dynamical equations must be solved at every time evaluation. Starting from the initial time t=t0t=t_{0}, the initial condition must satisfy the constraint (32) and the non-dynamical equations (36)-(39). We then evolve (ψ,P,Φ)(\psi,P,\Phi) to the next time step t=t1t=t_{1} with the dynamical equations (33)-(35). In order to obtain (α,β,a,Q)(\alpha,\beta,a,Q) at t=t1t=t_{1}, we integrate the non-dynamical equations (36)-(39) once more, now on the time slice t=t1t=t_{1}. At this point, the numerical accuracy of the solution may be gauged by checking how well the data on the time slice t=t1t=t_{1} satisfies the constraint (32). The numerical scheme can then be advanced to the next time step by repeating this procedure.

Unless stated otherwise, the initial field ψ\psi is given by an even Gaussian wave packet (defined in equation (47)) of height A=0.15A=0.15 and a width characterized by s=4s=4 positioned at r0=10r_{0}=10. An outer boundary rbr_{\mathrm{b}} of the system is placed far from the origin r=0r=0 and the initial wave packet, to prevent it from disturbing the collapsing process and to best simulate an asymptotically flat spacetime. In our case, rb=80r_{\mathrm{b}}=80 was used. Larger or slightly smaller values yield equivalent results. Space and time are uniformly discretized into intervals of size Δr\Delta r and Δt\Delta t, respectively. We decide here to keep the ratio Δt/Δr=0.2\Delta t/\Delta r=0.2 constant. A smaller ratio would also have been acceptable, but if Δt\Delta t is too large relative to Δr\Delta r, the system may become unstable.

To prevent any numerical instabilities from jeopardizing the computations, we also adopt two additional safety measures. Firstly — and most importantly — we use a series expansion near the boundaries to approximate the first and last integration points of the nondynamical fields (see appendix B). Secondly, we include Kreiss-Oliger dissipation terms [33]. These two measures serve to mitigate any possible emergent high-frequency modes near the boundaries and around the scalar field profile. As is standard, the artificial dissipation is chosen to be two orders higher than the convergence order of the spatial numerical derivatives.

The simulation was executed with several independently written codes, using different integration schemes, and all results matched. As an extra sanity check, the quantity 𝒞\mathcal{C}, defined in equation (40), was continuously and carefully surveilled to ensure it remained sufficiently small throughout the simulation. A more quantitative examination of 𝒞\mathcal{C} was also carried out, as figure 1 demonstrates. While the shape of 𝒞\mathcal{C} remains unchanged (left plot), its overall amplitude is inversely proportional to the square of the time step Δt\Delta t or, equivalently, the spatial gap Δr\Delta r (right plot). The quadratic convergence is expected. All codes — two of which are, in fact, the same codes as in [34, 35] and one of the two was also used in [36, 37] — exploit second-order methods to integrate the non-dynamical equations in space, and at least second-order methods for the integration in time. Explicitly, one used the trapezoidal method in space and the iterated Crank-Nicolson method with two iterations in time, while the other two used Heun’s method in space and higher order methods, mainly RK4, in time.

Refer to caption
Figure 1: The left plot displays the constraint 𝒞\mathcal{C}, as defined in equation (40), divided by the square of the time step Δt\Delta t, versus the radius rr, for two different values of Δt\Delta t. We reiterate that Δt=0.2Δr\Delta t=0.2\Delta r. As the shape remains unchanged for the two values of Δt\Delta t, we conclude that 𝒞(Δt)2\mathcal{C}\propto(\Delta t)^{2} and that the numerical convergence is of quadratic order as expected. The right plot portrays the average absolute value of 𝒞\mathcal{C} for 5r155\leq r\leq 15 (the region in the left plot for which 𝒞\mathcal{C} clearly is non-zero) against Δt\Delta t. A simple linear fit neatly illustrates the quadratic convergence order. The fit is here restricted to Δt103\Delta t\geq 10^{-3} values in order to avoid the small numerical noise that starts seeping in at lower step sizes (the excluded points are highlighted above). Both plots are generated at time t=2t=2. Albeit early after the numerical resolution is engaged, choosing different times tt does not affect the convergence order.

IV.2 Results

We now turn to the physical quantities thusly obtained444We remind the reader that as a consequence of having integrated out the shadowy mode, the integrated equations (32)-(39) are the same as in GR. Naturally, the results are then the same as in GR as far as the foliation-independent parts are concerned. Unlike in GR, however, the foliation-dependent parts are also physical in VCDM and thus will be investigated with care.. Figure 2 shows the areal radius Φ\Phi against the proper distance rr for some selected times tt. Notice that one of the displayed times, in this figure and the subsequent ones, is t14.1t\approx 14.1. As we shall explain in subsection IV.3, this corresponds to the AH formation time. At t=0t=0, we observe that Φr\Phi\approx r. As time proceeds, it only retains this behavior near r=0r=0 (in accordance with subsection III.4), falling and settling down into a plateau elsewhere. The height of this plateau, Φpl.0.65\Phi_{\text{pl.}}\approx 0.65, echoes in the other quantities, as the subsequent figures show. Henceforth, we shall use Φ\Phi as a radial coordinate to illustrate the evolution of all other quantities.

Refer to caption
Figure 2: The areal radius Φ\Phi plotted against the proper distance rr from the center at different times tt. One may observe the boundary conditions established in subsection III.4 in the shape of Φ\Phi. Following the fall of the right side of Φ\Phi, a plateau forms around Φpl.0.65\Phi_{\text{pl.}}\approx 0.65.

The other two dynamical quantities ψ\psi and PP are given for different times in figure 3. For the prescribed initial conditions, the scalar field ψ\psi starts by splitting into two parts. One moves out to r=r=\infty and vanishes, while the other falls to the origin r=0r=0 under its self-gravity and eventually settles near r=0r=0. Naturally, this foreshadows an AH formation (see subsection IV.3). A similar behavior is observed for the related variable PP (ψ\equiv\partial_{\perp}\psi, as defined in equation (11)).

Refer to caption
Figure 3: The scalar field ψ\psi (left) and PP (right) against the areal radius Φ\Phi for different values of time tt. The two sharp changes seen at t=25t=25 — the dip under zero for ψ\psi and the sharp turn for PP — are located around Φ=Φpl.\Phi=\Phi_{\text{pl.}}. The former is further detailed in the inset plot, where one can observe that the apparent spike is smooth. For the chosen range of Φ\Phi, the split of ψ\psi is not explicitly observable (as r0=10r_{0}=10).

In parallel, the results for the nondynamical fields are given in figure 4. One can observe how the lapse α\alpha and the shift β\beta collapse to zero below the threshold value of Φpl.\Phi_{\text{pl.}}. Far from the origin, an asymptotically flat value is manifestly recovered for both quantities. The variables aa and QQ simultaneously spike around the same threshold value of Φpl.\Phi_{\text{pl.}}.

Refer to caption
Figure 4: All four nondynamical fields α\alpha, aa, β\beta and QQ, here plotted against the areal radius Φ\Phi at different times tt. Similarly to figure 3, two inset plots for aa and QQ were added to explicitly show the peak smoothness.

IV.3 Apparent horizon formation

The quantity gΦΦg^{\Phi\Phi} (defined in equation (31)) is depicted for different times in the left plot of figure 5. As the behavior of ψ\psi already suggested (see figure 3), the collapse does in fact yield an AH, which occurs when gΦΦg^{\Phi\Phi} crosses zero. This condition is satisfied at t14.1t\approx 14.1, as the right plot of figure 5 shows. Furthermore, in figure 4, one can deduce that α\alpha and β\beta have not yet collapsed down to 0 at the time of the crossing, implying that the AH indeed forms before any singularity or breakdown of the time slicing does. Eventually, as the lapse and shift evolve towards zero inside the AH, gΦΦg^{\Phi\Phi} settles with a minimum around Φpl.0.65\Phi_{\text{pl.}}\approx 0.65 and an outermost crossing of zero at Φ0.88\Phi\approx 0.88. This constitutes the main achievement of this work, namely that the gravitational collapse of a massless scalar field in VCDM results in the manifestation of an AH before the breakdown of the time foliation. The final stage of the collapse is a static black hole.

Refer to caption
Figure 5: The left panel gives gΦΦg^{\Phi\Phi} with respect to the areal radius Φ\Phi at different times and demonstrates the creation of an AH (see equation (12)). At late time, gΦΦg^{\Phi\Phi} settles with a minimum around Φpl.\Phi_{\text{pl.}} and an outermost crossing near Φ0.88\Phi\approx 0.88. The right panel details the minimum of gΦΦg^{\Phi\Phi} as a function of time and portrays how the horizon forms at t14.1t\approx 14.1, at the “crossing”. A careful reader may wonder about the small cusp near t4.1t\approx 4.1: it is a consequence of ψ\psi splitting into two at the beginning of the simulation.

IV.4 Parameter variation

Until now, we have kept fixed the initial scalar field parameters of ψ0\psi_{0} as defined in equation (47) to A=0.15A=0.15 and s=4s=4. However, depending on the values assigned to these two parameters, an AH may or may not form. The parameter space where an AH does appear is illustrated in figure 6. A decrease in AA and/or an increase of ss may both “dilute” the scalar field enough to allow it to bounce back around the origin before any AH appears or the time slicing breaks down. An example of this is shown in the left plot of figure 7, where we have chosen s=8s=8 and plotted ψ\psi at different times throughout the evolution. The right plot of the same figure makes the non-formation of an AH evident: gΦΦg^{\Phi\Phi} never crosses zero.

Refer to caption
Figure 6: The minimal value (in time and space) that gΦΦg^{\Phi\Phi} reaches before t=30t=30, which, given that the initial wave packet of ψ\psi is placed at r0=10r_{0}=10, is enough time to determine whether a horizon forms or not (see equation (12)). The black cross indicates the A=0.15A=0.15 and s=4s=4 parameter pair that was previously used, while the black line approximately delimits min(gΦΦ)=0\min(g^{\Phi\Phi})=0. The range in AA was uniformly sampled in 301301 points from 0.00.0 to 0.30.3, while the values in ss were sampled in 176176 points uniformly spread from 2.02.0 to 9.09.0.
Refer to caption
Figure 7: Evolution of the scalar field ψ\psi for A=0.15A=0.15 and s=8s=8 on the left. Observe how the packet moves toward the origin before bouncing back. At t=25t=25, the bulk is mostly out of the represented range. The right plot gives min(gΦΦ)\min(g^{\Phi\Phi}) for the same configuration. Naturally, no AH forms (i.e. t:gΦΦ>0\forall t:\,g^{\Phi\Phi}>0).

V Summary and discussion

As is well-known, gravitational collapse in the framework of GR leads to the formation of a black hole, a trapped region of spacetime bounded by a marginally outer trapped surface, called the apparent horizon. Deviations from this picture may lead to contradictions through various observations such as GWs and black hole shadows, and so whether an alternative theory of gravity admits black holes or not is often used as a litmus test for its validity. It has been previously shown in [30] that a subset of the solutions of VCDM coincides exactly with solutions in GR when the 3+1-decomposition admits a constant trace of the extrinsic curvature. This subset of VCDM solutions consists of those for which the auxiliary field ϕ\phi and Lagrange multiplier λ\lambda are constant. Reassuringly, [31] shows that the spherical gravitational collapse of a homogeneous cloud of dust in VCDM coincides with a foliation of the Oppenheimer-Snyder collapse in GR for which indeed the aforementioned criteria are satisfied.

In the continuation of [31], we have numerically investigated the collapse of a massless scalar field in VCDM, in hopes of observing the formation of a black hole and whether its horizon appears before or after any singularity or foliation breakdown. Starting from a spherically symmetric ansatz of the metric and other relevant quantities in the total action (7), we derived the equations of motion. By requiring vacuum flat spacetime in the standard Minkowski time slicing at infinity and regularity at r=0r=0, we have shown ϕ\phi to be a constant and the trace of the extrinsic curvature to vanish. An immediate consequence, as implied by equation (26), was that λ\lambda also reduced to a constant. Hence, the sufficient conditions were satisfied for solutions of this system to coincide with ones in GR in the specified time foliation.

The equations of motion reduce to a constraint equation (32), three dynamical equations (33)-(35), and four non-dynamical ones (36)-(39). These were integrated with the boundary conditions in subsection III.4 and initial conditions in subsection III.5. Furthermore, the quantity 𝒞\mathcal{C} (equation (40)) was used for assessing the numerical accuracy of the integration. The formation of an AH was shown through gΦΦg^{\Phi\Phi} as defined in equation (31) and by checking the condition (12). To ensure numerical stability, the meshing in time and space was chosen such that Δt/Δr=0.2\Delta t/\Delta r=0.2. Moreover, emergent high-frequency oscillations were tamed by implementing Taylor expansions near the boundaries and Kreiss-Oliger dissipation terms.

First and foremost, as all three independent codes utilized second-order methods in space, it is reassuring that figure 1 confirms a quadratic convergence of the error. Secondly, as seen in figure 5, the solutions to equations (33)-(39) indeed lead to the formation of an AH. Concurrently considering figure 4, the lapse α\alpha and shift β\beta are everywhere non-zero at the time of the AH formation, meaning the time foliation preferred by the theory fully describes spacetime inside the black hole at the time of its formation. As the simulation proceeds, α\alpha and β\beta converge towards zero in a finite region inside — but not up to — the AH. The proper time in this region elapses slower and slower than that measured by an observer far away at infinity as the universe evolves, eventually standing still. In other words, a breakdown of the time foliation inside the AH is in the process, though from the viewpoint of observers outside the horizon, it will take infinitely long to actually happen. This is by all means not an issue, as there is also an infinite amount of time for the universe to evolve in VCDM. It does however imply that a UV-completion of the theory is needed to fully describe the inside of the black hole.

Further testing of VCDM is necessary to probe its full potential. As previously mentioned, [30] tested the corresponding Oppenheimer-Snyder case; this paper proceeded with studying a collapsing massless scalar field. We have observed the creation of a black hole, its AH forming prior to the breakdown of the time foliation and the formation of the singularity. In the future, one could consider a collapsing star made up of more realistic matter, such as a fluid, or adding further layers of properties, e.g. charge or angular momentum. While this would drastically complicate the dynamics of the collapse, it would certainly provide useful insight into the viability of VCDM.

Any form of non-symmetrical dynamics — rotating black holes, black hole and neutron star binaries, GWs — may prevent ϕ\phi from being constant in time. This would yield non-GR solutions of VCDM, which necessarily need to be studied if the theory is to be fully explored.

Acknowledgements.
P.M. acknowledges support from the Japanese Government (MEXT) scholarship for Research Student. A.F.J. acknowledges support from the Sweden Japan Foundation as well as from JASSO for international exchange students coming to Japan. The work of S.M. was supported in part by World Premier International Research Center Initiative, MEXT, Japan.

Appendix A Apparent horizon

In this appendix we derive the equations (12) and (13), starting from a spherically symmetric ansatz in the ADM formalism as the one in subsection III.1,

gμνdxμdxν=α2dt2+(dr+βdt)2+Φ2dΩ2,\displaystyle g_{\mu\nu}dx^{\mu}dx^{\nu}=-\alpha^{2}dt^{2}+(dr+\beta dt)^{2}+\Phi^{2}d\Omega^{2}\,, (50)

where α,β\alpha,\beta are positive functions of tt and rr, and Φ=Φ(t,r)\Phi=\Phi(t,r) is non-negative. Equation (13) can be found immediately by a redefinition of the radial coordinate rΦ(t,r)r\rightarrow\Phi(t,r):

gΦΦ=ΦxμΦxνgμν=(Φ)2+(rΦ)2,\displaystyle\begin{split}g^{\Phi\Phi}&=\dfrac{\partial\Phi}{\partial x^{\mu}}\dfrac{\partial\Phi}{\partial x^{\nu}}g^{\mu\nu}\\ &=-(\partial_{\perp}\Phi)^{2}+(\partial_{r}\Phi)^{2}\,,\end{split} (51)

where =(tβr)/α\partial_{\perp}=(\partial_{t}-\beta\partial_{r})/\alpha is the derivative in the direction perpendicular to the surfaces of the slicing. Although Φ\Phi will not be used as a radial coordinate in what follows, the above derivation will prove convenient shortly. Let us construct a pair of null vector fields kk and ll:

kμμ+r,\displaystyle k^{\mu}\partial_{\mu}\equiv\partial_{\perp}+\partial_{r}\,, (52)
lμμr.\displaystyle l^{\mu}\partial_{\mu}\equiv\partial_{\perp}-\partial_{r}\,. (53)

The vector field kk is tangent to a congruence of radially outgoing null curves, and similarly, ll is tangent to a congruence of radially ingoing ones. We acknowledge the relations

kμkμ=lμlμ=0,kμlμ=lμkμ=2.\displaystyle k^{\mu}k_{\mu}=l^{\mu}l_{\mu}=0,\quad k^{\mu}l_{\mu}=l^{\mu}k_{\mu}=-2\,. (54)

The part of gμνg^{\mu\nu} transverse to these two vector fields, which we will denote by σμν\sigma^{\mu\nu}, is given by

σμν=gμν+12(kμlν+lμkν).\displaystyle\sigma^{\mu\nu}=g^{\mu\nu}+\dfrac{1}{2}(k^{\mu}l^{\nu}+l^{\mu}k^{\nu})\,. (55)

This metric has components σtt=σtr=σrt=σrr=0\sigma^{tt}=\sigma^{tr}=\sigma^{rt}=\sigma^{rr}=0. The apparent horizon r=rAHr=r_{\text{AH}} is found by solving for a marginally outer trapped surface of the spacetime, that is, by finding when the expansion scalar θout\theta_{\text{out}} of the outgoing congruence, i.e. with tangent field kk, is 0. The expansion scalar θout\theta_{\text{out}} is given by the transverse projection of the covariant derivative of kk:

θoutσμννkμ=2Φ(Φ+rΦ).\displaystyle\theta_{\text{out}}\equiv\sigma^{\mu\nu}\nabla_{\nu}k_{\mu}=\dfrac{2}{\Phi}(\partial_{\perp}\Phi+\partial_{r}\Phi)\,. (56)

If we now set θout|r=rAH=0\theta_{\text{out}}|_{r=r_{\text{AH}}}=0, it follows that

(Φ+rΦ)|r=rAH=0.\displaystyle(\partial_{\perp}\Phi+\partial_{r}\Phi)\big{|}_{r=r_{\text{AH}}}=0\,. (57)

This implies the equality

gΦΦ|r=rAH=0,\displaystyle g^{\Phi\Phi}\big{|}_{r=r_{\text{AH}}}=0\,, (58)

where gΦΦg^{\Phi\Phi} is given in equation (51). Actually, the following identity can be easily derived:

gΦΦ=Φ24θinθout,\displaystyle g^{\Phi\Phi}=-\dfrac{\Phi^{2}}{4}\theta_{\text{in}}\theta_{\text{out}}\,, (59)

where

θinσμννlμ=2Φ(ΦrΦ),\displaystyle\theta_{\text{in}}\equiv\sigma^{\mu\nu}\nabla_{\nu}l_{\mu}=\dfrac{2}{\Phi}\left(\partial_{\perp}\Phi-\partial_{r}\Phi\right)\,, (60)

and one can numerically confirm that θin\theta_{\text{in}} does not vanish during the collapse. Consequently, if an apparent horizon exists, its position is given by a solution to equation (58). One can easily show that θin<0<θout\theta_{\text{in}}<0<\theta_{\text{out}} at the regular origin Φ=0\Phi=0 and in the asymptotically flat region Φ\Phi\to\infty. Therefore, the roots of gΦΦg^{\Phi\Phi} should be pairwise as long as they are not degenerate. The outer most root of gΦΦg^{\Phi\Phi} corresponds to the apparent horizon. It is then simply a matter of substitution using (11) to arrive at the right-hand side expression seen in equation (13).

Appendix B Taylor expansions near the boundaries

We will now discuss the Taylor expansions that were implemented near the boundaries to prevent high-frequency oscillations from appearing. This was only applied to the non-dynamical fields, as the dynamical fields were known in all of space at every time step.

Suppose we have NN dynamical fields fn(x)f_{n}(x) and MM non-dynamical fields gm(x)g_{m}(x), where n=1,,Nn=1,\ldots,N and m=1,,Mm=1,\ldots,M, and xx is the spatial variable discretized with step-length Δx\Delta x. Any quantity h(x)h(x) can be Taylor-expanded to some order kmaxk_{\rm max} around some point x0x_{0} accordingly,

h(x)=k=0kmaxh(k)(x0)k!(xx0)k+𝒪(xkmax+1).\displaystyle h(x)=\sum_{k=0}^{k_{\rm max}}\dfrac{h^{(k)}(x_{0})}{k!}(x-x_{0})^{k}+\mathcal{O}(x^{k_{\rm max}+1})\,. (61)

The following procedure is applicable to any point x0x_{0}.

  1. 1.

    Begin by Taylor expanding all quantities, both dynamical and non-dynamical, to a sufficiently high order kmaxk_{\rm max} around x0x_{0}. The order generally depends on how many points one wishes to manually set. If x0x_{0} is a boundary, then apply relevant boundary conditions by adjusting the coefficients adequately.

  2. 2.

    Suppose we are considering kmax+1k_{\rm max}+1 points xjx_{j} in the closest vicinity of x0x_{0} where j=0,1,,kmaxj=0,1,\ldots,k_{\rm max} (which includes x0x_{0}). Since the dynamical fields fnf_{n} are known at each point, their expansion coefficients fn(k)(x0)f_{n}^{(k)}(x_{0}) are easily computed by equating the given value of a dynamical field fnf_{n} at xjx_{j} with its respective Taylor expansion evaluated at the same point. This will yield a system of kmax+1k_{\rm max}+1 equations,

    {fn(xj)k=0kmaxfn(k)(x0)k!(xjx0)k:j=0,1,,kmax+1},\displaystyle\left\{f_{n}(x_{j})\approx\sum_{k=0}^{k_{\rm max}}\dfrac{f_{n}^{(k)}(x_{0})}{k!}(x_{j}-x_{0})^{k}:j=0,1,\dots,k_{\rm max}+1\right\}\,, (62)

    that can be solved for the kmax+1k_{\rm max}+1 unknown expansion coefficients fn(k)(x0)f_{n}^{(k)}(x_{0}) in terms of the known values fn(xj)f_{n}(x_{j}). Let us denote their thus computed value as f¯n(k)\overline{f}_{n}^{(k)}.

  3. 3.

    Insert the now fully numerical expressions f¯n(k)\overline{f}_{n}^{(k)} for the fn(k)(x0)f_{n}^{(k)}(x_{0}) into the Taylor expansions of fn(x)f_{n}(x):

    fn(x)=k=0kmaxf¯n(k)k!(xx0)k+𝒪(xkmax+1).\displaystyle f_{n}(x)=\sum_{k=0}^{k_{\rm max}}\dfrac{\overline{f}_{n}^{(k)}}{k!}(x-x_{0})^{k}+\mathcal{O}(x^{k_{\rm max}+1})\,. (63)
  4. 4.

    Insert all Taylor expansions fn(x)f_{n}(x) and gm(x)g_{m}(x) into the non-dynamical equations and gather together terms of each order. The coefficients in front of the terms will provide equations relating the non-dynamical expansion coefficients gm(k)(x0)g_{m}^{(k)}(x_{0}) to the fully numerical f¯n(k)\overline{f}_{n}^{(k)}. Solve these equations to acquire fully numerical solutions g¯m(k)\overline{g}_{m}^{(k)} for the non-dynamical expansion coefficients gm(k)(x0)g_{m}^{(k)}(x_{0}).

  5. 5.

    Insert the fully numerical solutions g¯m(k)\overline{g}_{m}^{(k)} for the expansion coefficients gm(k)(x0)g_{m}^{(k)}(x_{0}) into the Taylor expansions of the non-dynamical fields gm(x)g_{m}(x). We now have fully determined polynomial expressions for the non-dynamical fields around x0x_{0}:

    gm(x)k=0kmaxg¯m(k)k!(xx0)k.\displaystyle g_{m}(x)\approx\sum_{k=0}^{k_{\rm max}}\dfrac{\overline{g}_{m}^{(k)}}{k!}(x-x_{0})^{k}\,. (64)
  6. 6.

    Finally, manually set the value of gmg_{m} in any point close to x0x_{0} by simply evaluating its expansion in the said point.

In our case, this procedure was applied to the points r=0r=0 and r=rbr=r_{b}, where we set the values of the dynamical fields in the first three points and the last two points of the integration space [0,rb][0,r_{b}].

References