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

A hyperboloidal method for numerical simulations of multidimensional nonlinear wave equations

Oliver Rinne HTW Berlin – University of Applied Sciences, Faculty 4, Treskowallee 8, 10318 Berlin, Germany oliver.rinne@htw-berlin.de
Abstract

We consider the scalar wave equation with power nonlinearity in n+1n+1 dimensions. Unlike previous numerical studies, we go beyond the radial case and do not assume any symmetries for n=3n=3, and we only impose an SO(n1)(n-1) symmetry in higher dimensions. Our method is based on a hyperboloidal foliation of Minkowski spacetime and conformal compactification. We focus on the late-time power-law decay (tails) of the solutions and compute decay exponents for different spherical harmonic modes, for subcritical, critical and supercritical, focusing and defocusing nonlinear wave equations.

1 Introduction

This paper is concerned with the nonlinear wave equation (NLW)

Φ:=t2Φ+ΔΦ=μΦ|p1Φ,Φ:×n,\Box\Phi:=-\partial_{t}^{2}\Phi+\Delta\Phi=\mu\Phi|^{p-1}\Phi,\quad\Phi:\mathbb{R}\times\mathbb{R}^{n}\to\mathbb{R}, (1)

where p>1p>1 and μ=±1\mu=\pm 1, with μ=1\mu=-1 referred to as the focusing NLW and μ=1\mu=1 as the defocusing NLW.

This equation serves as a model for various nonlinear wavelike equations arising e.g. in fluid dynamics, optics, acoustics, plasma physics, general relativity and quantum field theory. Related equations include the nonlinear Schrödinger equation, the Korteweg-de Vries equation, the Klein-Gordon equation, Yang-Mills equations and wave map equations.

The NLW (1) is invariant under the rescaling

Φ(t,x)λ2p1Φ(λt,λx).\displaystyle\Phi(t,x)\to\lambda^{\frac{2}{p-1}}\Phi(\lambda t,\lambda x). (2)

The equation possesses a conserved energy

E(Φ,tΦ)=n(12(tΦ)2+12Φ2+μp+1|Φ|p+1)x.E(\Phi,\partial_{t}\Phi)=\int_{\mathbb{R}^{n}}\left(\textstyle\frac{1}{2}(\partial_{t}\Phi)^{2}+\textstyle\frac{1}{2}\|\nabla\Phi\|^{2}+\frac{\mu}{p+1}|\Phi|^{p+1}\right)\rmd x. (3)

The energy is invariant under the rescaling (2) iff

p=n+2n2=:pcrit,p=\frac{n+2}{n-2}=:p_{\mathrm{crit}}, (4)

in which case the NLW is called energy-critical. For p<pcritp<p_{\mathrm{crit}} the NLW is subcritical, for p>pcritp>p_{\mathrm{crit}} supercritical.

Solutions to the NLW can show rich behaviour due to the interplay of the dispersive wave operator and the nonlinear term. While small initial data will generally scatter, i.e. approach a solution to the linear wave equation, in the focusing case large initial data will cause blow-up of the solution. Futhermore, in the critical case stable finite-energy solitons exist that may prevent solutions from scattering.

In this paper we will mainly explore the late-time behaviour of solutions that scatter. In [1] it was proved using perturbative methods that in n=3n=3 spatial dimensions, spherically symmetric solutions to (1) decay as tp+1t^{-p+1} for p3p\geqslant 3. This is often referred to as a power-law tail. Here we will study such tails numerically beyond the spherically symmetric case.

In the standard approach to solving wavelike equations numerically, applied e.g. in [2, 3, 4, 5], the spatial domain is taken to be a large ball. Boundary conditions must be imposed at the surface of this ball in order to obtain a well-posed initial-boundary value problem. Typically a homogeneous Dirichlet boundary condition is used. This causes spurious reflections once the outgoing waves reach the boundary so that the numerical solution can only be trusted up to a certain time. Mapping the radial coordinate from (0,)(0,\infty) to a finite interval and discretising the compactified coordinate is not a reliable approach either because the wavelength w.r.t. the compactified coordinate decreases towards zero as the waves travel outwards, and they ultimately fail to be resolved numerically.

A different approach is to introduce a new time coordinate t~\tilde{t} such that the slices t~=const\tilde{t}=\mathrm{const} become asymptotically characteristic, e.g.

t~=ta2+r2\displaystyle\tilde{t}=t-\sqrt{a^{2}+r^{2}} (5)

with a constant aa. Such hyperboloidal slices approach future null infinity I+\mathrsfs{I}^{+}\,as rr\to\infty, the asymptotic region where outgoing null geodesics end. By means of a suitable compactification, I+\mathrsfs{I}^{+}\,can be brought to a finite radius on the numerical grid. No boundary conditions are needed there because all characteristics point towards the exterior of the domain, and the infinite “blueshift” of outgoing waves is eliminated.

The hyperboloidal method originated in general relativity, see [6] for a comprehensive review article. It has been applied to nonlinear wave equations in spherical symmetry [7], spherically symmetric linear scalar wave and Yang-Mills equations in fixed Schwarzschild spacetime [8] and coupled to the Einstein equations [9], and linear scalar wave equations without symmetries in Kerr spacetime [10].

As far as we are aware, the present paper is the first numerical study of the NLW (1) beyond spherical symmetry. We do not assume any symmetries in n=3n=3 spatial dimensions. In higher dimensions, we impose an SO(n1)(n-1) symmetry so that there is one effective angular coordinate.

Our numerical method is similar to that in [10, 11] in that it combines a finite-difference method in the radial direction with a spectral method in the angular directions. However, due to the nonlinear terms in our wave equation, we employ a pseudo-spectral collocation method.

This paper is organised as follows. In section 2 we derive the form of the wave equation in the framework of the hyperboloidal method. In section 3 we describe the numerical methods we use to solve this equation. We present our numerical results in section 4: a convergence test against exact linear solutions, a test of the energy balance on hyperboloidal slices, and power-law tails. The results are summarised and discussed in section 5. Exact solutions to the linear wave equation are derived in A.

2 Formulation

In this section we derive the form of the wave equation we will be solving numerically. In section 2.1 we foliate Minkowski spacetime into hyperboloidal slices of constant mean curvature and apply a conformal transformation to the metric. We work out the wave equation in terms of the conformally rescaled quantities in section 2.2 and write it in first-order in time, second-order in space form.

2.1 Hyperboloidal foliation of Minkowski spacetime

We start with the (n+1)(n+1)-dimensional Minkowski metric in spherical polar coordinates,

η=t2+r2+r2σ(n1),t,r(0,).\eta=-\rmd t^{2}+\rmd r^{2}+r^{2}\sigma^{(n-1)},\quad t\in\mathbb{R},\quad r\in(0,\infty). (6)

Here σ(n1)\sigma^{(n-1)} is the standard round metric on Sn1S^{n-1}, the (n1)(n-1)-dimensional unit sphere. Explicitly, we have

σ(1)=φ2,\displaystyle\sigma^{(1)}=\rmd\varphi^{2}, (7)
σ(2)=θ2+sin2θφ2,\displaystyle\sigma^{(2)}=\rmd\theta^{2}+\sin^{2}\theta\,\rmd\varphi^{2}, (8)

where θ[0,π]\theta\in[0,\pi] and φ[0,2π)\varphi\in[0,2\pi). More generally for n>3n>3,

σ(n1)=θ12+sin2θ1(θ22+sin2θ2((θn22+sin2θn2φ2)))\sigma^{(n-1)}=\rmd\theta_{1}^{2}+\sin^{2}\theta_{1}\left(\rmd\theta_{2}^{2}+\sin^{2}\theta_{2}(\ldots(\rmd\theta_{n-2}^{2}+\sin^{2}\theta_{n-2}\rmd\varphi^{2}))\right) (9)

with θ1,,θn2[0,π]\theta_{1},\ldots,\theta_{n-2}\in[0,\pi] and φ[0,2π)\varphi\in[0,2\pi).

We introduce a new time coordinate

t~=ta2+r2\tilde{t}=t-\sqrt{a^{2}+r^{2}} (10)

with a constant a=n/C>0a=n/C>0. The hypersurfaces t~=const\tilde{t}=\mathrm{const} are hyperboloids and CC is their constant mean curvature (up to a sign as discussed further below). In the new coordinates the Minkowski metric reads

η=t~22ra2+r2t~r+a2a2+r2r2+r2σ(n1).\eta=-\rmd\tilde{t}^{2}-\frac{2r}{\sqrt{a^{2}+r^{2}}}\,\rmd\tilde{t}\,\rmd r+\frac{a^{2}}{a^{2}+r^{2}}\,\rmd r^{2}+r^{2}\sigma^{(n-1)}. (11)

Let us also introduce a new radial coordinate r~\tilde{r} depending on rr only. We can arrange for the spatial metric to be conformally flat if we demand

a2a2+r2r2=r2r~2r~2rr~=ra2+r2r~a.\frac{a^{2}}{a^{2}+r^{2}}\rmd r^{2}=\frac{r^{2}}{\tilde{r}^{2}}\rmd\tilde{r}^{2}\;\Leftrightarrow\;\frac{\rmd r}{\rmd\tilde{r}}=\frac{r\sqrt{a^{2}+r^{2}}}{\tilde{r}a}. (12)

This ordinary differential equation can be solved explicitly. If we impose the boundary condition r~1\tilde{r}\to 1 as rr\to\infty, we obtain

r=2ar~1r~2=2nr~C(1r~2).r=\frac{2a\tilde{r}}{1-\tilde{r}^{2}}=\frac{2n\tilde{r}}{C(1-\tilde{r}^{2})}. (13)

From (10) we observe that on a slice t~=const\tilde{t}=\mathrm{const}, as the conformal boundary r~1\tilde{r}\to 1 is approached, we have

t,r,trt~=const,t,r\to\infty,\quad t-r\to\tilde{t}=\mathrm{const}, (14)

so the slice becomes asymptotically characteristic and r~=1\tilde{r}=1 corresponds to I+\mathrsfs{I}^{+}\,.

The metric now takes the form

η\displaystyle\eta =\displaystyle= t~22r2r~at~r~+r2r~2(r~2+r~2σ(n1))\displaystyle-\rmd\tilde{t}^{2}-\frac{2r^{2}}{\tilde{r}a}\rmd\tilde{t}\,\rmd\tilde{r}+\frac{r^{2}}{\tilde{r}^{2}}\left(\rmd\tilde{r}^{2}+\tilde{r}^{2}\sigma^{(n-1)}\right) (15)
=\displaystyle= Ω2[α~2t~2+(r~+β~r~t~)2+r~2σ(n1)]=:Ω2η~.\displaystyle\Omega^{-2}\left[-\tilde{\alpha}^{2}\rmd\tilde{t}^{2}+(\rmd\tilde{r}+\tilde{\beta}^{\tilde{r}}\rmd\tilde{t})^{2}+\tilde{r}^{2}\sigma^{(n-1)}\right]=:\Omega^{-2}\tilde{\eta}.

Here we have introduced the conformal factor

Ω=r~r=C2n(1r~2),\Omega=\frac{\tilde{r}}{r}=\frac{C}{2n}(1-\tilde{r}^{2}), (16)

the conformal shift vector β~\tilde{\beta}, which only has a radial component

β~r~=r~a=Cnr~,\tilde{\beta}^{\tilde{r}}=-\frac{\tilde{r}}{a}=-\frac{C}{n}\tilde{r}, (17)

and the conformal lapse function

α~=C2n(1+r~2)\tilde{\alpha}=\frac{C}{2n}(1+\tilde{r}^{2}) (18)

so that α~2(β~r~)2=Ω2\tilde{\alpha}^{2}-(\tilde{\beta}^{\tilde{r}})^{2}=\Omega^{2}.

From now on we will work entirely with the conformal metric η~\tilde{\eta}. Its induced metric γ~\tilde{\gamma} on the t~=const\tilde{t}=\mathrm{const} slices is flat,

γ~=r~2+r~2σ(n1).\tilde{\gamma}=\rmd\tilde{r}^{2}+\tilde{r}^{2}\sigma^{(n-1)}. (19)

The conformal extrinsic curvature of the t~=const\tilde{t}=\mathrm{const} slices is111 We use the sign convention ν~γ~ab=2K~ab\mathcal{L}_{\tilde{\nu}}\tilde{\gamma}_{ab}=-2\tilde{K}_{ab}. For the t~=const\tilde{t}=\mathrm{const} slices to intersect I+\mathrsfs{I}^{+}\,we need the constant C>0C>0 so that the shift (17) is negative at r~=1\tilde{r}=1. Spacetime indices a,ba,b are raised and lowered using the conformal metric η~\tilde{\eta}, and repeated indices are summed over.

K~ab=K~nγ~ab\tilde{K}_{ab}=\frac{\tilde{K}}{n}\tilde{\gamma}_{ab} (20)

with the conformal mean curvature given by

K~=α~1C=2nr~2+1.\tilde{K}=-\tilde{\alpha}^{-1}C=-\frac{2n}{\tilde{r}^{2}+1}. (21)

Its Lie derivative along the conformal future-directed timelike unit normal ν~\tilde{\nu} to the t~=const\tilde{t}=\mathrm{const} slices is

ν~K~=α~1β~r~r~K~=8nr~2(r~2+1)3.\mathcal{L}_{\tilde{\nu}}\tilde{K}=-\tilde{\alpha}^{-1}\tilde{\beta}^{\tilde{r}}\partial_{\tilde{r}}\tilde{K}=\frac{8n\tilde{r}^{2}}{(\tilde{r}^{2}+1)^{3}}. (22)

Combining the contracted Gauss and Ricci equations, we can relate the Ricci scalar R~\tilde{R} of η~\tilde{\eta} to the Ricci scalar R~γ~\tilde{R}_{\tilde{\gamma}} of γ~\tilde{\gamma}:

R~=R~γ~2α~1Δ~α~2ν~K~+K~2+K~abK~ab.\tilde{R}=\tilde{R}_{\tilde{\gamma}}-2\tilde{\alpha}^{-1}\tilde{\Delta}\tilde{\alpha}-2\mathcal{L}_{\tilde{\nu}}\tilde{K}+\tilde{K}^{2}+\tilde{K}_{ab}\tilde{K}^{ab}. (23)

Here

Δ~=r~2+(n1)r~1r~+r~2Δ̊(n1)\tilde{\Delta}=\partial_{\tilde{r}}^{2}+(n-1){\tilde{r}}^{-1}\partial_{\tilde{r}}+\tilde{r}^{-2}\mathring{\Delta}^{(n-1)} (24)

is the standard flat Laplacian of the flat conformal metric γ\gamma and Δ̊(n1)\mathring{\Delta}^{(n-1)} is the Laplace-Beltrami operator on the sphere Sn1S^{n-1}. Explicitly, we have

Δ̊(1)=φ2,\displaystyle\mathring{\Delta}^{(1)}=\partial_{\varphi}^{2}, (25)
Δ̊(2)=θ2+cotθθ+1sin2θφ2.\displaystyle\mathring{\Delta}^{(2)}=\partial_{\theta}^{2}+\cot\theta\,\partial_{\theta}+\frac{1}{\sin^{2}\theta}\partial_{\varphi}^{2}. (26)

In our numerical implementation for higher dimensions n>3n>3, we shall impose an SO(n1)(n-1) symmetry so that all but one of the angular coordinates are suppressed. The angular Laplacian then reduces to

Δ̊(n1)=θ2+(n2)cotθθ.\mathring{\Delta}^{(n-1)}=\partial_{\theta}^{2}+(n-2)\cot\theta\,\partial_{\theta}. (27)

Using the fact that R~γ~=0\tilde{R}_{\tilde{\gamma}}=0 and the form of the conformal extrinsic curvature (20), we obtain from (23)

R~=2α~1Δ~α~2ν~K~+n+1nK~2=4n[r~4+(n5)r~2+n](r~2+1)3.\tilde{R}=-2\tilde{\alpha}^{-1}\tilde{\Delta}\tilde{\alpha}-2\mathcal{L}_{\tilde{\nu}}\tilde{K}+\frac{n+1}{n}\tilde{K}^{2}=\frac{4n[-\tilde{r}^{4}+(n-5)\tilde{r}^{2}+n]}{(\tilde{r}^{2}+1)^{3}}. (28)

2.2 Wave equation

The scalar field Φ\Phi is supposed to obey the NLW

Φ=μ|Φ|p1Φ,\Box\Phi=\mu|\Phi|^{p-1}\Phi, (29)

where \Box is the d’Alembertian of the Minkowski metric η\eta.

Defining a conformally rescaled scalar field

Φ~=Ω(1n)/2Φ\tilde{\Phi}=\Omega^{(1-n)/2}\Phi (30)

and using the conformal transformation properties of the d’Alembertian and the Ricci scalar, we have

~Φ~n14nR~Φ~=Ω(n+3)/2(Φn14nRΦ)=μΩ[p(n1)n3]/2|Φ~|p1Φ~,\tilde{\Box}\tilde{\Phi}-\frac{n-1}{4n}\tilde{R}\tilde{\Phi}=\Omega^{-(n+3)/2}\left(\Box\Phi-\frac{n-1}{4n}R\Phi\right)=\mu\Omega^{[p(n-1)-n-3]/2}|\tilde{\Phi}|^{p-1}\tilde{\Phi}, (31)

where on the right-hand side we have used (29) and the fact that η\eta is flat, R=0R=0. This equation is regular at I+\mathrsfs{I}^{+}\,, where Ω=0\Omega=0, iff n2n\geqslant 2 and

p(n1)n30pn+3n1=:pconf.p(n-1)-n-3\geqslant 0\;\Leftrightarrow\;p\geqslant\frac{n+3}{n-1}=:p_{\mathrm{conf}}. (32)

For any n3n\geqslant 3 we have pconf<pcritp_{\mathrm{conf}}<p_{\mathrm{crit}}, where pcritp_{\mathrm{crit}} was defined in (4), cf. table 1. Thus the conformal approach is capable of treating both energy-subcritical and supercritical NLWs.

n234567pconf537/329/510/6pcrit537/329/5\begin{array}[]{c||c|c|c|c|c|c|c}n&2&3&4&5&6&7&\cdots\\ \hline\cr p_{\mathrm{conf}}&5&3&7/3&2&9/5&10/6&\cdots\\ \hline\cr p_{\mathrm{crit}}&\infty&5&3&7/3&2&9/5&\cdots\end{array}
Table 1: The smallest allowed exponent pconfp_{\mathrm{conf}} of the nonlinearity for the conformal approach to be applicable, as compared with the energy-critical exponent pcritp_{\mathrm{crit}}, depending on the number of space dimensions nn. For n=2n=2 the NLW is energy-subcritical for any p>1p>1.

Performing an (n+1)(n+1)-decomposition of the conformal d’Alembertian, (31) becomes

ν~2Φ~+Δ~Φ~+K~ν~Φ~+α~1α~,iΦ~,in14nR~Φ~=μΩ[p(n1)n3]/2|Φ~|p1Φ~,-\mathcal{L}_{\tilde{\nu}}^{2}\tilde{\Phi}+\tilde{\Delta}\tilde{\Phi}+\tilde{K}\mathcal{L}_{\tilde{\nu}}\tilde{\Phi}+\tilde{\alpha}^{-1}\tilde{\alpha}^{,i}\tilde{\Phi}_{,i}-\frac{n-1}{4n}\tilde{R}\tilde{\Phi}=\mu\Omega^{[p(n-1)-n-3]/2}|\tilde{\Phi}|^{p-1}\tilde{\Phi}, (33)

where spatial indices ii are raised and lowered using the flat conformal metric γ~\tilde{\gamma}. We can write this equation in first-order in time form by introducing an auxiliary field Π~\tilde{\Pi},

ν~Φ~\displaystyle\mathcal{L}_{\tilde{\nu}}\tilde{\Phi} =:\displaystyle=: Π~,\displaystyle\tilde{\Pi}, (34)
ν~Π~\displaystyle\mathcal{L}_{\tilde{\nu}}\tilde{\Pi} =\displaystyle= Δ~Φ~+K~Π~+α~1α~,iΦ~,in14nR~Φ~μΩ[p(n1)n3]/2|Φ~|p1Φ~.\displaystyle\tilde{\Delta}\tilde{\Phi}+\tilde{K}\tilde{\Pi}+\tilde{\alpha}^{-1}\tilde{\alpha}^{,i}\tilde{\Phi}_{,i}-\frac{n-1}{4n}\tilde{R}\tilde{\Phi}-\mu\Omega^{[p(n-1)-n-3]/2}|\tilde{\Phi}|^{p-1}\tilde{\Phi}. (35)

Writing out the Lie derivative ν~=α~1(t~β~)\mathcal{L}_{\tilde{\nu}}=\tilde{\alpha}^{-1}(\partial_{\tilde{t}}-\mathcal{L}_{\tilde{\beta}}), we arrive at

Φ~,t~=β~r~Φ~,r~+α~Π~,\displaystyle\tilde{\Phi}_{,\tilde{t}}=\tilde{\beta}^{\tilde{r}}\tilde{\Phi}_{,\tilde{r}}+\tilde{\alpha}\tilde{\Pi}, (36)
Π~,t~=r~1n[r~n1(β~r~Π~+α~Φ~,r~)],r~+α~r~2Δ̊(n2)Φ~n14nα~R~Φ~\displaystyle\tilde{\Pi}_{,\tilde{t}}=\tilde{r}^{1-n}\left[\tilde{r}^{n-1}\left(\tilde{\beta}^{\tilde{r}}\tilde{\Pi}+\tilde{\alpha}\tilde{\Phi}_{,\tilde{r}}\right)\right]_{,\tilde{r}}+\tilde{\alpha}\tilde{r}^{-2}\mathring{\Delta}^{(n-2)}\tilde{\Phi}-\frac{n-1}{4n}\tilde{\alpha}\tilde{R}\tilde{\Phi}
μα~Ω[p(n1)n3]/2|Φ~|p1Φ~.\displaystyle\qquad-\mu\tilde{\alpha}\Omega^{[p(n-1)-n-3]/2}|\tilde{\Phi}|^{p-1}\tilde{\Phi}. (37)

We have double-checked using computer algebra that (36)–(37) are equivalent to (29) if the coordinate transformation given by (10) and (13) is applied directly.

Equations (36)–(37) are perfectly regular at future null infinity r~=1\tilde{r}=1, where Ω=0\Omega=0, provided that ppconfp\geqslant p_{\mathrm{conf}} (or μ=0\mu=0). No boundary conditions are needed there because all characteristics point towards the exterior of the domain.

It should be noted that we have written the radial principal part of (37) in conservative form. The finite differencing (section 3.1) will be applied in precisely this order. This was found to be essential for the numerical evolutions to be stable.

3 Numerical methods

In this section we describe the numerical methods we use in order to solve the NLW (36)–(37). We begin with the spatial discretisation in section 3.1. Next we explain how the equation is integrated forward in time in section 3.2, and we address how to eliminate high-frequency instabilities. Finally we provide details on how spatial integrals needed for various diagnostics are computed in section 3.3. The code has been written in Python using libraries such as NumPy [12] and SciPy [13].

3.1 Spatial discretisation

We use a hybrid method for the spatial discretisation, which combines a finite-difference method in the radial direction with a pseudo-spectral method in the angular directions. The latter has the advantage that it is easier to build in parity conditions by an appropriate choice of expansion functions, as will be explained below. Furthermore, if we introduce a logically rectangular grid with equidistant grid points in the standard spherical coordinates θ\theta and φ\varphi on the two-sphere, the distance between neighbouring φ\varphi-grid points becomes very small near the poles. This leads to a severe restriction on the time step due to the Courant-Friedrichs-Lewy (CFL) condition if a finite-difference method is used. With a pseudo-spectral method, it is straightforward to remove an increasing number of higher (more oscillatory) φ\varphi-modes as the axis is approached (section 3.2), thus allowing for larger time steps.

We take the spatial dimension to be n=3n=3 without symmetries first. Higher dimensions with SO(n1)\mathrm{SO}(n-1) symmetry will be treated as a special case.

We introduce equidistant grid points in each dimension,

r~i=(i+12)hr~,i=0,1,,Nr~1,hr~=1/(Nr~12),\displaystyle\tilde{r}_{i}=(i+\textstyle\frac{1}{2})h_{\tilde{r}},\quad\;i=0,1,\ldots,N_{\tilde{r}}-1,\quad h_{\tilde{r}}=1/(N_{\tilde{r}}-\textstyle\frac{1}{2}), (38)
θj=(j+12)hθ,j=0,1,,Nθ1,hθ=π/Nθ,\displaystyle\theta_{j}=(j+\textstyle\frac{1}{2})h_{\theta},\quad j=0,1,\ldots,N_{\theta}-1,\quad h_{\theta}=\pi/N_{\theta}, (39)
φk=khφ,k=0,1,,Nφ1,hφ=2π/Nφ.\displaystyle\varphi_{k}=kh_{\varphi},\qquad\;\;k=0,1,\ldots,N_{\varphi}-1,\quad h_{\varphi}=2\pi/N_{\varphi}. (40)

The radial grid is staggered at the origin, whereas the outermost grid point coincides with I+\mathrsfs{I}^{+}\,(r~=1)\tilde{r}=1). The grid in the elevation θ\theta is staggered about the axis. The reason for this choice is that there are terms in the wave equation that are formally singular at the origin (r~=0\tilde{r}=0) and on the axis (sinθ=0\sin\theta=0), which would be difficult to evaluate if there were grid points at the origin and on the axis.

Fourth-order finite differences are used in order to approximate the radial derivatives. In the interior a centred stencil is used. Setting ui:=u(r~i)u_{i}:=u(\tilde{r}_{i}) for a smooth function uu and leaving out its angular dependence for simplicity,

ui(ui28ui1+8ui+1ui+2)/(12hr~),\displaystyle u^{\prime}_{i}\approx(u_{i-2}-8u_{i-1}+8u_{i+1}-u_{i+2})/(12h_{\tilde{r}}), (41)
ui′′(ui2+16ui130ui+16ui+1ui+2)/(12hr~2)\displaystyle u^{\prime\prime}_{i}\approx(-u_{i-2}+16u_{i-1}-30u_{i}+16u_{i+1}-u_{i+2})/(12h_{\tilde{r}}^{2}) (42)

for i=0,1,,Nr~3i=0,1,\ldots,N_{\tilde{r}}-3, where \approx stands for equality up to a truncation error of the order O(hr~4)O(h_{\tilde{r}}^{4}). Evaluating the above stencils at i=0i=0 and i=1i=1 requires values of uu at the ghost points r~1=hr~/2\tilde{r}_{-1}=-h_{\tilde{r}}/2 and r~2=3hr~/2\tilde{r}_{-2}=-3h_{\tilde{r}}/2, which we copy from interior points according to

u(r~,θ,φ)=u(r~,πθ,π+φ),u(-\tilde{r},\theta,\varphi)=u(\tilde{r},\pi-\theta,\pi+\varphi), (43)

satisfied by any smooth function uu. We require NφN_{\varphi} to be even, which ensures that if there is a grid point at φ\varphi, there is also a grid point at π+φ\pi+\varphi.

Near the outer boundary at r~=1\tilde{r}=1, one-sided differences are used,

uNr~2(uNr~5+6uNr~418uNr~3+10uNr~2+3uNr~1)/(12hr~),\displaystyle u^{\prime}_{N_{\tilde{r}}-2}\approx(-u_{N_{\tilde{r}}-5}+6u_{N_{\tilde{r}}-4}-18u_{N_{\tilde{r}}-3}+10u_{N_{\tilde{r}}-2}+3u_{N_{\tilde{r}}-1})/(12h_{\tilde{r}}), (44)
uNr~1(3uNr~516uNr~4+36uNr~348uNr~2+25uNr~1)/(12hr~),\displaystyle u^{\prime}_{N_{\tilde{r}}-1}\approx(3u_{N_{\tilde{r}}-5}-16u_{N_{\tilde{r}}-4}+36u_{N_{\tilde{r}}-3}-48u_{N_{\tilde{r}}-2}+25u_{N_{\tilde{r}}-1})/(12h_{\tilde{r}}), (45)
uNr~2′′(uNr~66uNr~5+14uNr~44uNr~315uNr~2+10uNr~1)/(12hr~2),\displaystyle u^{\prime\prime}_{N_{\tilde{r}}-2}\approx(u_{N_{\tilde{r}}-6}-6u_{N_{\tilde{r}}-5}+14u_{N_{\tilde{r}}-4}-4u_{N_{\tilde{r}}-3}-15u_{N_{\tilde{r}}-2}+10u_{N_{\tilde{r}}-1})/(12h_{\tilde{r}}^{2}), (46)
uNr~1′′(10uNr~6+61uNr~5156uNr~4+214uNr~3154uNr~2+45uNr~1)\displaystyle u^{\prime\prime}_{N_{\tilde{r}}-1}\approx(-10u_{N_{\tilde{r}}-6}+61u_{N_{\tilde{r}}-5}-156u_{N_{\tilde{r}}-4}+214u_{N_{\tilde{r}}-3}-154u_{N_{\tilde{r}}-2}+45u_{N_{\tilde{r}}-1})
/(12hr~2),\displaystyle\qquad\qquad/(12h_{\tilde{r}}^{2}), (47)

the truncation error again being O(hr~4)O(h_{\tilde{r}}^{4}).

Our pseudo-spectral method is based on Fourier expansions in both θ\theta and φ\varphi. This allows us to employ fast Fourier transform (FFT) techniques [14], which would not be available if we expanded the fields in spherical harmonics.

Leaving out its radial dependence for simplicity, a smooth function uu is expanded as

u(θ,φ)l=0Nθ1(cos(lθ)m=0mevenNφ/21almmφ+sin(lθ)m=1moddNφ/21almmφ).u(\theta,\varphi)\approx\sum_{l=0}^{N_{\theta}-1}\left(\cos(l\theta)\sum_{m=0\atop{m\,\mathrm{even}}}^{N_{\varphi}/2-1}a_{lm}\,\rme^{\rmi m\varphi}+\sin(l\theta)\sum_{m=1\atop{m\,\mathrm{odd}}}^{N_{\varphi}/2-1}a_{lm}\,\rme^{\rmi m\varphi}\right). (48)

Any smooth function uu on the two-sphere must obey

u(θ,φ)=u(θ,π+φ),u(π+θ,φ)=u(πθ,π+φ),u(-\theta,\varphi)=u(\theta,\pi+\varphi),\qquad u(\pi+\theta,\varphi)=u(\pi-\theta,\pi+\varphi), (49)

and the way we expand the even and odd mm-modes separately in (48) enforces this. For uu to be real, the coefficients alma_{lm} in (48) must be complex, and we may use

Re(almmφ)=Realmcos(mφ)Imalmsin(mφ).\mathrm{Re}(a_{lm}\rme^{\rmi m\varphi})=\mathrm{Re}\,a_{lm}\cos(m\varphi)-\mathrm{Im}\,a_{lm}\sin(m\varphi). (50)

In order to compute the expansion coefficients alma_{lm} from the point values ujk:=u(θj,φk)u_{jk}:=u(\theta_{j},\varphi_{k}), we first perform a real FFT w.r.t. φ\varphi using scipy.fft.rfft. Next, we perform a discrete cosine transform w.r.t. θ\theta for the even-mm modes (scipy.fft.dct) and a discrete sine transform for the odd-mm modes (scipy.fft.dst). The (default) Type II versions of the discrete cosine and sine transforms implemented in scipy are appropriate for our staggered θ\theta grid.

Derivatives of the expansion can now be computed analytically by simply differentiating the basis functions cos(lθ)\cos(l\theta), sin(lθ)\sin(l\theta) and mφ\rme^{\rmi m\varphi}. Finally in order to obtain the values of the derivatives at the grid points (θj,φk)(\theta_{j},\varphi_{k}), we apply the inverse discrete cosine and sine transforms w.r.t. θ\theta, followed by an inverse real FFT w.r.t. φ\varphi.

In the code the main object is the array of values of the unknowns Φ~\tilde{\Phi} and Π~\tilde{\Pi} at the grid points (r~i,θj,φk)(\tilde{r}_{i},\theta_{j},\varphi_{k}). Terms in the evolution equations, especially nonlinear terms, are evaluated and added pointwise.

We remark that a purely spectral (rather than pseudo-spectral collocation) method such as the one used in [10] cannot be employed in our case because of the nonlinear term in the wave equation we consider.

nn dimensions with SO(n1)\mathrm{SO}(n-1) symmetry.

In this case there is no φ\varphi-dependence and the expansion (48) is replaced with

u(θ)l=0Nθ1alcos(lθ)u(\theta)\approx\sum_{l=0}^{N_{\theta}-1}a_{l}\cos(l\theta) (51)

with real coefficients ala_{l}. This satisfies the symmetry conditions

u(θ)=u(θ),u(π+θ)=u(πθ).u(-\theta)=u(\theta),\qquad u(\pi+\theta)=u(\pi-\theta). (52)

3.2 Time stepping and filtering

We employ the method of lines, whereby the evolution equations (36)–(37) are first discretised in space as described in the previous subsection, leading to a system of ordinary differential equations (ODEs), one for each grid point. These ODEs are then integrated forward in time using the standard fourth-order Runge-Kutta method.

In order to obtain a stable finite-difference method, fifth-order Kreiss-Oliger dissipation [15] is applied in the radial direction, whereby the term

(Qu)i=ε64hr~(ui36ui2+15ui120ui+15ui+16ui+2+ui+3)\displaystyle(Qu)_{i}=\frac{\varepsilon}{64h_{\tilde{r}}}(u_{i-3}-6u_{i-2}+15u_{i-1}-20u_{i}+15u_{i+1}-6u_{i+2}+u_{i+3})
=ε64hr~5(6ur~6)i+O(hr~7)\displaystyle\qquad\;=\frac{\varepsilon}{64}h_{\tilde{r}}^{5}\left(\frac{\rmd^{6}u}{\rmd\tilde{r}^{6}}\right)_{i}+\Or(h_{\tilde{r}}^{7}) (53)

for i=0,1,,Nr~4i=0,1,\ldots,N_{\tilde{r}}-4 is added to the right-hand side of the discretised evolution equations for both Φ~\tilde{\Phi} and Π~\tilde{\Pi}. Ghost points near the origin are filled according to (43). No dissipation is added at the outermost three radial grid points. We typically take the parameter ε=0.2\varepsilon=0.2. It should be noted that the artificial dissipation term (53) is of higher order in hr~h_{\tilde{r}} than the truncation error of the finite-difference method.

In a pseudo-spectral method, nonlinear terms such as the one in our wave equation can cause high frequencies to be represented incorrectly due to aliasing [16]. We address this by applying spectral filtering according to the 2/3 rule, whereby the top third of all θ\theta-modes are removed at each substep of the Runge-Kutta algorithm.

In order to combat the clustering of grid points near the poles of the two-sphere, we filter out an increasing number of the highest Fourier modes w.r.t. φ\varphi as θ\theta approaches 0 or π\pi. Specifically, we remove the proportion 1sinθ1-\sin\theta of all φ\varphi-Fourier modes (beginning at the highest frequencies) [17].

The time step is restricted by the smallest distance between neighbouring grid points, and this occurs between neighbouring θ\theta-grid points at the smallest radius: Δxmin=r~0hθ\Delta x_{\min}=\tilde{r}_{0}h_{\theta}, where r~0=hr~/2\tilde{r}_{0}=h_{\tilde{r}}/2. The time step is then taken to be Δt~=λΔxmin\Delta\tilde{t}=\lambda\Delta x_{\min}, where 0<λ<10<\lambda<1 according to the CFL condition. We typically choose λ=0.8\lambda=0.8.

3.3 Spatial integration

For expressions such as the energy and flux (section 4.2) we will need to compute integrals on a t~=const\tilde{t}=\mathrm{const} slice and on its spherical boundary. The radial integration is performed using Simpson’s rule (taking Nr~N_{\tilde{r}} to be even),

01u(r~)r~hr~3(278u0+178u1+4u2+2u3+4u4+2u5+\displaystyle\int_{0}^{1}u(\tilde{r})\rmd\tilde{r}\approx\frac{h_{\tilde{r}}}{3}\left(\frac{27}{8}u_{0}+\frac{17}{8}u_{1}+4u_{2}+2u_{3}+4u_{4}+2u_{5}+\ldots\right.
+4uNr~4+2uNr~3+4uNr~2+uNr~1),\displaystyle\left.+4u_{N_{\tilde{r}}-4}+2u_{N_{\tilde{r}}-3}+4u_{N_{\tilde{r}}-2}+u_{N_{\tilde{r}}-1}\right), (54)

where the error is of the same order as the truncation error of the finite-difference scheme, O(hr~4)\Or(h_{\tilde{r}}^{4}). The modified coefficients of u0u_{0} and u1u_{1} are due to the fact that the grid is staggered at the origin, with a first grid point at r~0=hr~/2\tilde{r}_{0}=h_{\tilde{r}}/2, and uu is assumed to be an even function of r~\tilde{r} here.

The spectral expansion on the right-hand side of (48) can easily be integrated exactly:

S2u𝑑S(2)=02π0πu(θ,φ)sinθθφ2πl=0levenNθ12al01l2.\int_{S^{2}}u\,dS^{(2)}=\int_{0}^{2\pi}\int_{0}^{\pi}u(\theta,\varphi)\,\sin\theta\,\rmd\theta\,\rmd\varphi\approx 2\pi\sum_{l=0\atop{l\,\mathrm{even}}}^{N_{\theta}-1}\frac{2a_{l0}}{1-l^{2}}. (55)

nn dimensions with SO(n1)\mathrm{SO}(n-1) symmetry.

In this case (51) integrates to

Sn1u𝑑S(n1)=Sn20πu(θ)sinn2θθdS(n2)A(n2)l=0Nθ1alIn2,l,\int_{S^{n-1}}u\,dS^{(n-1)}=\int_{S^{n-2}}\int_{0}^{\pi}u(\theta)\sin^{n-2}\theta\,\rmd\theta\,dS^{(n-2)}\approx A^{(n-2)}\sum_{l=0}^{N_{\theta}-1}a_{l}\,I_{n-2,l}, (56)

where

A(n2)=2π(n1)/2Γ(n12)A^{(n-2)}=\frac{2\pi^{(n-1)/2}}{\Gamma(\textstyle\frac{n-1}{2})} (57)

is the area of Sn2S^{n-2}, and the integrals

In2,l:=0πsinn2θcos(lθ)θI_{n-2,l}:=\int_{0}^{\pi}\sin^{n-2}\theta\,\cos(l\theta)\,\rmd\theta (58)

are computed numerically to high precision using scipy.integrate.quad in the initialisation phase of the code.

4 Results

In this section we present our numerical results. We first perform a convergence test against exact solutions to the linear wave equation in section 4.1. Next we derive the energy balance that the scalar field must obey on hyperboloidal slices, and we check it is satisfied in our numerical evolutions (section 4.2). Finally in section 4.3 we investigate the late-time power-law decay (tails) of the solutions.

4.1 Convergence test against an exact linear solution

Given the complexity of the algebraic operations involved in deriving the evolution equations, it is very useful to have an exact solution at hand in order to check that the implementation is correct, and that the numerical method converges as expected.

In A we construct exact solutions to the linear wave equation. They are based on an expansion in spherical harmonics and are given in terms of a mode function, which we take to be

F(x)=Axexp[12(xσ)2].F(x)=A\,x\,\exp\left[-\frac{1}{2}\left(\frac{x}{\sigma}\right)^{2}\right]. (59)

We choose A=1A=1 and σ=1\sigma=1.

For the convergence test in the case of n=3n=3 without symmetries, we consider a superposition of two modes with spherical harmonic indices (l,m)=(1,1)(l,m)=(1,1) and (2,2)(2,-2). In the case of n=5n=5 with SO(4)(4) symmetry, we take two modes with l=1l=1 and 22.

The initial data are set according to Φ~0:=Φ~(t~=0)=Φ~exact(t~=t~0)\tilde{\Phi}_{0}:=\tilde{\Phi}(\tilde{t}=0)=\tilde{\Phi}_{\mathrm{exact}}(\tilde{t}=\tilde{t}_{0}) and similarly for Π~\tilde{\Pi}, where we choose t~0=15\tilde{t}_{0}=-15 so that the solution is initially ingoing and centred roughly in the middle of the conformal radial interval [0,1][0,1].

We now evolve these initial data and compare with the exact solution. In figure 1 we plot the L2L^{2} norm of the error

Φ~Φ~exactL2=[01r~n1r~Sn1(Φ~Φ~exact)2𝑑S(n1)]1/2\|\tilde{\Phi}-\tilde{\Phi}_{\mathrm{exact}}\|_{L^{2}}=\left[\int_{0}^{1}\tilde{r}^{n-1}\rmd\tilde{r}\int_{S^{n-1}}(\tilde{\Phi}-\tilde{\Phi}_{\mathrm{exact}})^{2}\,dS^{(n-1)}\right]^{1/2} (60)

as a function of time for three different radial resolutions at fixed angular resolution. We observe that doubling the resolution decreases the error nearly by a factor of 24=162^{4}=16, as expected for a fourth-order accurate finite-difference method.

We remark that at the chosen angular resolution, the spherical harmonics are represented exactly by our pseudo-spectral method, so there is no discretisation error in the angular dimensions for these linear evolutions.

Refer to caption
Refer to caption
Figure 1: Convergence test against an exact linear solution. Left panel: n=3n=3 dimensions without symmetries, right panel: n=5n=5 with SO(4)(4) symmetry. Shown is the L2L^{2} norm of the error of Φ~\tilde{\Phi} w.r.t. the exact linear solution as a function of time for three different radial resolutions, from top to bottom: Nr~=250N_{\tilde{r}}=250 (blue), Nr~=500N_{\tilde{r}}=500 (red) and Nr~=1000N_{\tilde{r}}=1000 (green).

4.2 Energy balance

In this section we derive and verify numerically the energy balance that the scalar field obeys on our hyperboloidal foliation. On standard slices of Minkowski spacetime approaching spacelike infinity, the energy (3) is conserved. In contrast, the energy computed on hyperboloidal slices decreases during the evolution. This is due to the fact that hyperboloidal slices intersect I+\mathrsfs{I}^{+}\,, so that outgoing waves pass through this conformal boundary, carrying away energy. We will compute this energy flux explicitly. A similar energy balance was derived in [10] in the context of linear scalar fields in Kerr spacetime, where an additional inner boundary arises at the black hole horizon.

The scalar field Φ\Phi is associated with an energy-momentum tensor

Tab=aΦbΦ12ηabcΦcΦμp+1|Φ|p+1ηab,T_{ab}=\nabla_{a}\Phi\nabla_{b}\Phi-\textstyle\frac{1}{2}\eta_{ab}\nabla_{c}\Phi\nabla^{c}\Phi-\frac{\mu}{p+1}|\Phi|^{p+1}\eta_{ab}, (61)

where \nabla is the covariant derivative of the Minkowski metric η\eta. The vanishing of its divergence, bTab=0\nabla^{b}T_{ab}=0, is equivalent with the NLW (29). The Killing vector k:=/t=/t~k:=\partial/\partial t=\partial/\partial\tilde{t} gives rise to the conserved current

Ea=Takbb=TatE^{a}=T^{a}{}_{b}k^{b}=T^{a}{}_{t} (62)

satisfying

aEa=0.\nabla_{a}E^{a}=0. (63)

Let Σt~\Sigma_{\tilde{t}} denote the slice of constant hyperboloidal time t~\tilde{t} with future-directed unit normal ν\nu (normalised w.r.t. η\eta) and induced Riemannian metric γab=ηab+νaνb\gamma_{ab}=\eta_{ab}+\nu_{a}\nu_{b}. Furthermore let ss denote the outward-directed normal to the timelike surface r=constr=\mathrm{const} (again normalised w.r.t. η\eta) and hab=ηabsasbh_{ab}=\eta_{ab}-s_{a}s_{b} the induced pseudo-Riemannian metric on such a surface. Applying Gauss’ theorem to (63), the energy balance reads

E(t~2)E(t~1):=Σt~2νaEadetγrn1θΣt~1νaEadetγrn1θ\displaystyle E(\tilde{t}_{2})-E(\tilde{t}_{1}):=\int_{\Sigma_{\tilde{t}_{2}}}\nu_{a}E^{a}\sqrt{\det\gamma}\,\rmd r\,\rmd^{n-1}\theta-\int_{\Sigma_{\tilde{t}_{1}}}\nu_{a}E^{a}\sqrt{\det\gamma}\,\rmd r\,\rmd^{n-1}\theta (64)
=t~1t~2Sn1saEadet|h|n1θt~|I+=:F(t~1,t~2),\displaystyle\qquad=\int_{\tilde{t}_{1}}^{\tilde{t}_{2}}\int_{S^{n-1}}s_{a}E^{a}\sqrt{\det|h|}\,\rmd^{n-1}{\bf\theta}\,\rmd\tilde{t}\,\Big{|}_{\mathrsfs{I}^{+}}=:F(\tilde{t}_{1},\tilde{t}_{2}), (65)

where θ\theta stands collectively for the n1n-1 angular coordinates on Sn1S^{n-1}.

Expressing Φ\Phi in terms of the conformally rescaled scalar field Φ~\tilde{\Phi} defined in 30, we obtain for the energy

E(t~)=C4n01r~n1r~Sn1dS(n1){(1+r~2)[Π~2+Φ~,r~2+1r~2̊(n1)Φ~2\displaystyle E(\tilde{t})=\frac{C}{4n}\int_{0}^{1}\tilde{r}^{n-1}\rmd\tilde{r}\int_{S^{n-1}}dS^{(n-1)}\left\{(1+\tilde{r}^{2})\left[\tilde{\Pi}^{2}+\tilde{\Phi}_{,\tilde{r}}^{2}+\frac{1}{\tilde{r}^{2}}\|\mathring{\nabla}^{(n-1)}\tilde{\Phi}\|^{2}\right.\right.
+2μΩ[p(n1)n3]/21p+1|Φ~|p+1]\displaystyle\left.\qquad+2\mu\Omega^{[p(n-1)-n-3]/2}\frac{1}{p+1}|\tilde{\Phi}|^{p+1}\right]
+2(n1)r~21r~2+1r~Φ~,r~Φ~4rΦ~,r~Π~+(n1)2r~2r~2+1Φ~2}.\displaystyle\qquad\left.+2(n-1)\frac{\tilde{r}^{2}-1}{\tilde{r}^{2}+1}\,\tilde{r}\tilde{\Phi}_{,\tilde{r}}\tilde{\Phi}-4r\tilde{\Phi}_{,\tilde{r}}\tilde{\Pi}+(n-1)^{2}\frac{\tilde{r}^{2}}{\tilde{r}^{2}+1}\tilde{\Phi}^{2}\right\}. (66)

Here ̊(n1)\mathring{\nabla}^{(n-1)} refers to the gradient on Sn1S^{n-1} and dS(n1)=detσ(n1)n1θdS^{(n-1)}=\sqrt{\det\sigma^{(n-1)}}\rmd^{n-1}\theta to its area element, e.g. for n=3n=3

̊(2)Φ2=Φ,θ2+1sin2θΦ,φ2,dS(2)=sinθθφ.\|\mathring{\nabla}^{(2)}\Phi\|^{2}=\Phi_{,\theta}^{2}+\frac{1}{\sin^{2}\theta}\,\Phi_{,\varphi}^{2},\qquad dS^{(2)}=\sin\theta\,\rmd\theta\,\rmd\varphi. (67)

In the first line of (4.2), we recognize the usual energy expression for the linear scalar field. The second line of (4.2) contains the contribution from the nonlinear potential, and the terms in the third line arise from the conformal rescaling.

The flux integral takes the form

F(t~1,t~2)=C2n2t~1t~2t~S(n1)𝑑S(n1)(Φ~,r~Π~)2|r~=1.F(\tilde{t}_{1},\tilde{t}_{2})=-\frac{C^{2}}{n^{2}}\int_{\tilde{t}_{1}}^{\tilde{t}_{2}}\rmd\tilde{t}\int_{S^{(n-1)}}dS^{(n-1)}(\tilde{\Phi}_{,\tilde{r}}-\tilde{\Pi})^{2}\,\Big{|}_{\tilde{r}=1}. (68)

It is is manifestly negative, reflecting the fact that the waves carry away energy as they pass through I+\mathrsfs{I}^{+}\,.

We now verify the energy balance numerically for two cases: n=3,p=5n=3,\,p=5 (critical) and n=5,p=3n=5,\,p=3 (supercritical). In both cases we impose an SO(n1)(n-1) symmetry. The initial data are taken to be momentarily static with

Φ~0=Aexp[(r~r~0σ)2]Yl(θ),\tilde{\Phi}_{0}=A\exp\left[-\left(\frac{\tilde{r}-\tilde{r}_{0}}{\sigma}\right)^{2}\right]\,Y_{l}(\theta), (69)

and Φ~,t=0\tilde{\Phi}_{,t}=0 at t=0t=0 implies

Π~0=2r~1+r~2Φ~0,r~.\tilde{\Pi}_{0}=\frac{2\tilde{r}}{1+\tilde{r}^{2}}\tilde{\Phi}_{0,\tilde{r}}. (70)

We superpose two modes with l=2l=2 and l=3l=3 and choose the parameters r~0=0.3\tilde{r}_{0}=0.3 and σ=0.07\sigma=0.07. The amplitude for both modes is taken to be A=12A=12 for n=3,p=5n=3,\,p=5 and A=200A=200 for n=5,p=3n=5,\,p=3, which is smaller than but close to the critical amplitude beyond which blow-up occurs in the focusing case (μ=1\mu=-1). The numerical resolution is taken to be Nr~=4000,Nθ=16N_{\tilde{r}}=4000,\,N_{\theta}=16.

Figure 2 demonstrates that the energy balance

E(t~)F(0,t~)=E(0)E(\tilde{t})-F(0,\tilde{t})=E(0) (71)

is well satisfied numerically, i.e., the sum of the numerically computed terms on the left-hand side is indeed nearly constant. For an evolution with a defocusing nonlinearity (μ=1\mu=1), the initial energy is higher than for a focusing nonlinearity (μ=1\mu=-1) due to the different sign of the potential energy term in (4.2).

Refer to caption
Refer to caption
Figure 2: Energy balance for the nonlinear wave equation. Left panel: n=3,p=5n=3,\,p=5, right panel: n=5,p=3n=5,\,p=3, both with SO(n1)(n-1) symmetry. Shown is the energy E(t~)E(\tilde{t}) (blue, monotonically decreasing), the integrated flux F(0,t~)-F(0,\tilde{t}) (red, monotonically increasing) and their sum (green, nearly constant). Solid lines correspond to an evolution with a focusing nonlinearity (μ=1\mu=-1), dashed lines to a defocusing nonlinearity (μ=1\mu=1).

We also compute this potential energy contribution alone,

Epot(t~)=Cμ2n01r~n1r~Sn1𝑑S(n1)(1+r~2)Ω[p(n1)n3]/21p+1|Φ~|p+1E_{\mathrm{pot}}(\tilde{t})=\frac{C\mu}{2n}\int_{0}^{1}\tilde{r}^{n-1}\rmd\tilde{r}\int_{S^{n-1}}dS^{(n-1)}(1+\tilde{r}^{2})\Omega^{[p(n-1)-n-3]/2}\frac{1}{p+1}|\tilde{\Phi}|^{p+1} (72)

and its ratio to the total energy, Epot(t~)/E(t~)E_{\mathrm{pot}}(\tilde{t})/E(\tilde{t}). The result is shown in figure 3 and indicates that the potential energy contribution becomes negligible at late times.

Refer to caption
Refer to caption
Figure 3: Ratio Epot(t~)/E(t~)E_{\mathrm{pot}}(\tilde{t})/E(\tilde{t}) of the potential energy to the total energy as a function of time for the same evolutions as in figure 2. Left panel: n=3,p=5n=3,\,p=5, right panel: n=5,p=3n=5,\,p=3, both with SO(n1)(n-1) symmetry. Solid lines correspond to an evolution with a focusing nonlinearity (μ=1\mu=-1), dashed lines to a defocusing nonlinearity (μ=1\mu=1).

4.3 Tails

In this section we focus on the late-time behaviour of the scalar field, the so-called power-law tails. We extract the scalar field at various radii and decompose it into spherical harmonics. Each mode shows a characteristic power-law decay t~q\sim\tilde{t}^{-q} at late times. We investigate how the decay exponent qq depends on the power pp of the nonlinearity, the spherical harmonic mode indices (l,m)(l,m) and the extraction radius.

At a fixed extraction radius r~ex\tilde{r}_{\mathrm{ex}}, we expand the scalar field in spherical harmonics, here in n=3n=3 dimensions:

Φ~|r~=r~ex=l=0m=llΦ~lm(t~)Ylm(θ,φ).\tilde{\Phi}|_{\tilde{r}=\tilde{r}_{\mathrm{ex}}}=\sum_{l=0}^{\infty}\sum_{m=-l}^{l}\tilde{\Phi}_{lm}(\tilde{t})\,Y_{lm}(\theta,\varphi). (73)

For each of the modes we form the local power index

qlm(t~):=lnΦ~lmlnt~|r~=r~ex.q_{lm}(\tilde{t}):=-\frac{\rmd\ln\tilde{\Phi}_{lm}}{\rmd\ln\tilde{t}}\Big{|}_{\tilde{r}=\tilde{r}_{\mathrm{ex}}}. (74)

If this approaches a constant qlmq_{lm} as t~\tilde{t}\to\infty then the mode decays asymptotically as Φ~lmt~qlm\tilde{\Phi}_{lm}\sim\tilde{t}^{-q_{lm}}.

Our first observation is that the asymptotic decay rate appears to be independent of the azimuthal spherical harmonic index mm in three spatial dimensions. In order to illustrate this, figure 4 shows an evolution of static initial data (69)-(70) containing two modes with (l,m)=(2,1)(l,m)=(2,1) and (2,2)(2,2). Via the nonlinearity (p=5p=5, focusing), modes with (l,m)=(0,0)(l,m)=(0,0) and (2,0)(2,0) are also excited during the evolution. The l=2l=2 modes all decay at the same rate, q2m=6q_{2m}=6 irrespective of mm.

Refer to caption
Refer to caption
Figure 4: Modes Φ~lm\tilde{\Phi}_{lm} (left panel) and local power indices qlmq_{lm} (right panel) extracted at r~ex=0.5\tilde{r}_{\mathrm{ex}}=0.5 for an evolution in n=3n=3 dimensions with a focusing p=5p=5 nonlinearity. Static initial data containing two modes with l=2,m=1l=2,\,m=1 (amplitude A=6A=6) and l=2,m=2l=2,\,m=2 (amplitude A=12A=12) are chosen. The remaining initial data parameters are r~0=0.3\tilde{r}_{0}=0.3 and σ=0.07\sigma=0.07 for both modes. The numerical resolution is Nr~=2000,Nθ=Nφ=8N_{\tilde{r}}=2000,\,N_{\theta}=N_{\varphi}=8.

In the following we therefore focus on the m=0m=0 case only, i.e. we impose axisymmetry for n=3n=3, and more generally an SO(n1)(n-1) symmetry in nn dimensions.

In figure 5 we demonstrate convergence of the numerically computed local power index as the radial resolution is increased. In the following we will always choose the highest resolution shown here, Nr~=4000N_{\tilde{r}}=4000 and Nθ=12N_{\theta}=12. We have checked that increasing the angular resolution has a negligible effect on the results.

Refer to caption
Refer to caption
Figure 5: Convergence test for the local power index in n=3n=3 dimensions under axisymmetry with a p=5p=5 focusing nonlinearity. Shown is the local power index q2q_{2} of the l=2l=2 mode extracted at a finite radius (left panel) and at I+\mathrsfs{I}^{+}\,(right panel). Static initial data containing two modes with l=2l=2 and l=3l=3 are chosen, both with A=12,r~0=0.3A=12,\,\tilde{r}_{0}=0.3 and σ=0.07\sigma=0.07. The radial resolutions are Nr~=1000N_{\tilde{r}}=1000 (dotted), Nr~=2000N_{\tilde{r}}=2000 (dashed) and Nr~=4000N_{\tilde{r}}=4000 (solid). The angular resolution is fixed at Nθ=12N_{\theta}=12.

Next we compare the time evolution of the local power index at different extraction radii in figure 6. We observe that the local power index of a given mode approaches the same constant at all finite extraction radii, but in in general a different constant at I+\mathrsfs{I}^{+}\,(r~=1\tilde{r}=1).

Refer to caption
Figure 6: Dependence of the local power index on the extraction radius rexr_{\mathrm{ex}} in n=3n=3 dimensions under axisymmetry with a p=5p=5 focusing nonlinearity. Shown is the local power index q2q_{2} of the l=2l=2 mode extracted at four different radii, from top to bottom: r~ex=0.5\tilde{r}_{\mathrm{ex}}=0.5 (blue), 0.90.9 (red), 0.990.99 (blue) and 11 (black). Static initial data containing two modes with l=2l=2 and l=3l=3 are chosen, both with A=12,r~0=0.3A=12,\,\tilde{r}_{0}=0.3 and σ=0.07\sigma=0.07. The numerical resolution is Nr~=4000,Nθ=12N_{\tilde{r}}=4000,\,N_{\theta}=12.

Finally we investigate the dependence of the decay rates on the spherical harmonic index ll. We show results for two selected cases, n=3,p=5n=3,\,p=5, focusing (figure 7) and n=5,p=3n=5,\,p=3, defocusing (figure 8). In the n=5n=5 case the higher-ll modes are very small. We had to use higher precision (longdouble) for this run. Due to the smallness of the modes, computation of the local power index via (74) becomes numerically unstable. Instead, we have performed a power-law fit to the modes Φ~l\tilde{\Phi}_{l} shown in figure 8 in order to obtain the decay rates.

Refer to caption
Refer to caption
Figure 7: Time evolution of the local power indices qlq_{l} in n=3n=3 dimensions under axisymmetry with a p=5p=5 focusing nonlinearity, extracted at a finite radius (left panel) and at I+\mathrsfs{I}^{+}\,(right panel). From bottom to top in the left panel: l=0l=0 (black), 11 (blue), 22 (red) and 33 (green). In the right panel the curves for l=0,1,2l=0,1,2 almost coincide. Static initial data containing two modes with l=2l=2 and l=3l=3 are chosen, both with A=12,r~0=0.3A=12,\,\tilde{r}_{0}=0.3 and σ=0.07\sigma=0.07. The numerical resolution is Nr~=4000,Nθ=12N_{\tilde{r}}=4000,\,N_{\theta}=12.
Refer to caption
Refer to caption
Figure 8: Time evolution of the modes Φ~l\tilde{\Phi}_{l} in n=5n=5 dimensions under SO(4)(4) symmetry with a p=3p=3 defocusing nonlinearity, extracted at a finite radius (left panel) and at I+\mathrsfs{I}^{+}\,(right panel). From top to bottom: l=0l=0 (black), 11 (blue), 22 (red) and 33 (green). Static initial data containing two modes with l=2l=2 and l=3l=3 are chosen, both with A=20,r~0=0.3A=20,\,\tilde{r}_{0}=0.3 and σ=0.07\sigma=0.07. The numerical resolution is Nr~=4000,Nθ=12N_{\tilde{r}}=4000,\,N_{\theta}=12.

Table 2 summarises the asymptotic decay rates qlq_{l} found in our numerical evolutions for n=3n=3 and n=5n=5 and various values of pp (subcritical and supercritical). We have compared focusing and defocusing evolutions and find the same asymptotic decay rates in both cases. In addition to the static initial data (69)–(70) used in the evolutions shown here, we have also tried different initial data taken from the exact linear solutions derived in A. The observed decay rates are identical. The static data have the advantage that they have an initially outgoing component so the tail forms earlier and with a higher amplitude than for the data from the linear solutions, which are initially ingoing.

n=3p=3p=4p=5p=6p=7l=02|13|24|35|46|5l=14|24|25|36|47|5l=26|36|36|37|48|5l=38|48|48|48?|49?|5?\begin{array}[]{c||c|c|c|c|c}n=3&p=3&p=4&p=5&p=6&p=7\\ \hline\cr l=0&2|1&3|2&4|3&5|4&6|5\\ \hline\cr l=1&4|2&4|2&5|3&6|4&7|5\\ \hline\cr l=2&6|3&6|3&6|3&7|4&8|5\\ \hline\cr l=3&8|4&8|4&8|4&8?|4&9?|5?\end{array}
n=5p=2p=3l=04|25|3l=16|36|3l=28|48?|4l=310|510?|5\begin{array}[]{c||c|c}n=5&p=2&p=3\\ \hline\cr l=0&4|2&5|3\\ \hline\cr l=1&6|3&6|3\\ \hline\cr l=2&8|4&8?|4\\ \hline\cr l=3&10|5&10?|5\end{array}
Table 2: Asymptotic decay rates qlq_{l} for n=3n=3 and n=5n=5 under SO(n1)(n-1) symmetry, for various values of the power pp of the nonlinearity, determined from numerical evolutions. For each ll, the values of qlq_{l} at a finite extraction radius (left) and at I+\mathrsfs{I}^{+}\,(right) are shown. Values marked ? are uncertain.

5 Conclusion and discussion

This paper is concerned with the nonlinear wave equation

Φ:=t2Φ+ΔΦ=μΦ|p1Φ,Φ:×n\Box\Phi:=-\partial_{t}^{2}\Phi+\Delta\Phi=\mu\Phi|^{p-1}\Phi,\quad\Phi:\mathbb{R}\times\mathbb{R}^{n}\to\mathbb{R} (75)

with p>1p>1, μ=±1\mu=\pm 1 in n3n\geqslant 3 spatial dimensions. Unlike previous studies, we do not assume radial symmetry. In n=3n=3 we do not assume any symmetries, and in higher dimensions we impose an SO(n1)(n-1) symmetry so that there is one effective angular coordinate.

We introduce a foliation of Minkowski spacetime into hyperboloidal slices of constant mean curvature, combined with a conformal compactification. This avoids the need for artificial timelike boundaries and allows us to numerically construct the solution in the entire future of the initial hyperboloidal slice, all the way up to future null infinity I+\mathrsfs{I}^{+}\,, where future-directed null gedodesics end.

Our numerical approach combines a fourth-order finite difference method in the radial coordinate with a Fourier pseudo-spectral method in the angular coordinate(s). We construct exact solutions to the linear wave equation and demonstrate fourth-order convergence of our code as the radial resolution is increased.

Unlike on standard slices approaching spatial infinity, the energy of the scalar field on hyperboloidal slices is not conserved. It decreases due to the energy flux at I+\mathrsfs{I}^{+}\,. We derive this energy balance and show that it is well satisfied in our numerical evolutions.

The main result of this paper concerns the late-time power-law decay of the solutions. We expand the evolved field into spherical harmonics and determine decay exponents of the individual modes for various values of the exponent pp of the nonlinearity (subcritical, critical and supercritical). The decay exponents appear to be the same for focusing and defocusing nonlinearities (μ=±1\mu=\pm 1). They are independent of the extraction radius as long as this radius is finite, but at I+\mathrsfs{I}^{+}\,different (smaller) decay rates are found. In three dimensions the decay exponents are found to be independent of the azimuthal spherical harmonic index mm, which justifies our assumption of axisymmetry or more generally SO(n1)(n-1) symmetry. Our results in dimensions n=3n=3 and n=5n=5 are summarised in table 2.

At a finite radius, we would expect the scalar field to show the same decay in standard coordinates t,rt,r as in hyperboloidal coordinates t~,r~\tilde{t},\tilde{r}. This leads us to the following

Conjecture 1

Consider the nonlinear wave equation (75) in n=3n=3 spatial dimensions with pp\in\mathbb{N}, p3p\geqslant 3. Expanding the field in spherical harmonics

Φ(t,r,θ,φ)=l=0m=llΦlm(t,r)Ylm(θ,φ),\Phi(t,r,\theta,\varphi)=\sum_{l=0}^{\infty}\sum_{m=-l}^{l}\Phi_{lm}(t,r)\,Y_{lm}(\theta,\varphi), (76)

the modes decay asymptotically (as tt\to\infty) as

Φlm(t,r)tql,ql=max(l+p1, 2l+2)\Phi_{lm}(t,r)\sim t^{-q_{l}},\qquad q_{l}=\max(l+p-1,\,2l+2) (77)

at any fixed finite radius rr. On slices Σt~\Sigma_{\tilde{t}} approaching future null infinity I+\mathrsfs{I}^{+}\,, where t,rt,r\to\infty, trt~t-r\to\tilde{t} and the conformal factor Ω(r)=O(r1)\Omega(r)=\Or(r^{-1}), the conformally rescaled scalar field Φ~=Ω(r)(1n)/2Φ\tilde{\Phi}=\Omega(r)^{(1-n)/2}\,\Phi decays at I+\mathrsfs{I}^{+}\,asymptotically as

Φ~lm(t~)t~q~l,q~l=max(p2,l+1).\tilde{\Phi}_{lm}(\tilde{t})\sim\tilde{t}^{-\tilde{q}_{l}},\qquad\tilde{q}_{l}=\max(p-2,\,l+1). (78)

Indeed, this is consistent with [1], where the authors proved using perturbative methods that in n=3n=3 spatial dimensions under spherical symmetry (l=0l=0) and p3p\geqslant 3 the solution decays asymptotically as tp+1t^{-p+1} at a finite radius. It would be very interesting to try and prove the above conjecture for l>0l>0.

In dimension n=5n=5, a similar form of the decay exponents consistent with table 2 would be

ql=max(l+p+2, 2l+4),q~l=max(p,l+2)q_{l}=\max(l+p+2,\,2l+4),\qquad\tilde{q}_{l}=\max(p,\,l+2) (79)

for the decay at a finite radius and at I+\mathrsfs{I}^{+}\,, respectively. More data would be needed to corroborate such a conjecture for n=5n=5 though. This proves difficult numerically because the field decays very rapidly for higher exponents pp of the nonlinearity. Higher than longdouble precision would be needed.

Apart from tails, there are various other aspects of nonlinear wave equations that could be investigated using our hyperboloidal evolution code, in particular the nature of singularity formation (blow-up) and the threshold between blow-up and scattering. For example, this threshold was investigated numerically for the focusing cubic (p=3p=3) wave equation in the subcritical dimension n=3n=3 in radial symmetry in [7] (see also a similar study for the nonlinear Klein-Gordon equation [4]). In [18] the focusing cubic wave equation was considered in the supercritical dimensions n=5n=5 and n=7n=7 and again threshold solutions were identified. It will be interesting to analyse these situations beyond radial symmetry.

Appendix A Exact solutions, initial data

Here we work out exact solutions to the linear wave equation, which are used to test the convergence of the code in section 4.1. We first assume dimension n=3n=3 without symmetries in A.1; higher dimensions will be treated in A.2. The solutions are constructed in standard spherical polar coordinates first; finally in A.3 they are transformed to hyperboloidal coordinates.

A.1 n=3n=3 without symmetries

The solution to the linear wave equation

Φ=0\Box\Phi=0 (80)

in standard spherical polar coordinates can be decomposed into spherical harmonics,

Φ=l=0m=llΦlm(t,r)Ylm(θ,φ).\Phi=\sum_{l=0}^{\infty}\sum_{m=-l}^{l}\Phi_{lm}(t,r)\,Y_{lm}(\theta,\varphi). (81)

The spherical harmonics are eigenfunctions of the Laplace-Beltrami operator on S2S^{2},

Δ̊(2)Ylm(θ,φ)=l(l+1)Ylm(θ,φ).\mathring{\Delta}^{(2)}\,Y_{lm}(\theta,\varphi)=-l(l+1)Y_{lm}(\theta,\varphi). (82)

They have the form

Ylm(θ,φ)=NlmPlm(θ)mφ,Y_{lm}(\theta,\varphi)=N_{lm}P_{lm}(\theta)\,\rme^{\rmi m\varphi}, (83)

where Plm(θ)P_{lm}(\theta) are the associated Legendre functions and NlmN_{lm} are normalisation constants chosen such that

S2YlmYlm𝑑S(2)=02π0πYlm(θ,φ)Ylm(θ,φ)sinθθφ=δllδmm.\int_{S^{2}}Y_{lm}Y_{l^{\prime}m^{\prime}}dS^{(2)}=\int_{0}^{2\pi}\int_{0}^{\pi}Y_{lm}(\theta,\varphi)Y_{l^{\prime}m^{\prime}}^{*}(\theta,\varphi)\sin\theta\,\rmd\theta\,\rmd\varphi=\delta_{ll^{\prime}}\delta_{mm^{\prime}}. (84)

The (complex) spherical harmonics are implemented in scipy.special.sph_harm_y. A suitable real basis of spherical harmonics is obtained by taking

{2(1)mImYl|m|,m<0,Yl0,m=0,2(1)mReYlm,m>0.\left\{\begin{array}[]{ll}\sqrt{2}(-1)^{m}\mathrm{Im}Y_{l|m|},&m<0,\vskip 3.0pt plus 1.0pt minus 1.0pt\\ Y_{l0},&m=0,\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \sqrt{2}(-1)^{m}\mathrm{Re}Y_{lm},&m>0.\end{array}\right. (85)

The wave equation implies that the radial functions Φlm(t,r)\Phi_{lm}(t,r) in (81) obey an equation of Euler-Poisson-Darboux type,

Φlm,tt+Φlm,rr+2rΦlm,rl(l+1)r2Φlm=0.-\Phi_{lm,tt}+\Phi_{lm,rr}+\frac{2}{r}\Phi_{lm,r}-\frac{l(l+1)}{r^{2}}\Phi_{lm}=0. (86)

Since this equation does not depend on mm explicitly, we leave out the index mm on Φlm\Phi_{lm} in the following. Solutions have the form

Φl±(t,r)=k=0lckrk1F(lk)(r±t),\Phi_{l}^{\pm}(t,r)=\sum_{k=0}^{l}c_{k}\,r^{-k-1}F^{(l-k)}(r\pm t), (87)

where FF is an arbitrary mode function, and - refers to an outgoing and ++ to an ingoing solution. The coefficients ckc_{k} can be determined by recursion. We provide explicit solutions for l=0,1,2l=0,1,2:

Φ0±(t,r)\displaystyle\Phi_{0}^{\pm}(t,r) =\displaystyle= r1F(r±t),\displaystyle r^{-1}F(r\pm t), (88)
Φ1±(t,r)\displaystyle\Phi_{1}^{\pm}(t,r) =\displaystyle= r1F(r±t)r2F(r±t),\displaystyle r^{-1}F^{\prime}(r\pm t)-r^{-2}F(r\pm t), (89)
Φ2±(t,r)\displaystyle\Phi_{2}^{\pm}(t,r) =\displaystyle= r1F′′(r±t)3r2F(r±t)+3r3F(r±t).\displaystyle r^{-1}F^{\prime\prime}(r\pm t)-3r^{-2}F^{\prime}(r\pm t)+3r^{-3}F(r\pm t). (90)

We observe that Φl=O(r1)\Phi_{l}^{-}=\Or(r^{-1}) as I+\mathrsfs{I}^{+}\,is approached (rr\to\infty, rtr-t finite).

If we take the mode function FF to be odd and form the linear combination Φl++Φl\Phi_{l}^{+}+\Phi_{l}^{-}, the result turns out to be regular at the origin r=0r=0.

A.2 n3n\geqslant 3 with SO(n1)\mathrm{SO}(n-1) symmetry

In this case we expand

Φ=l=0Φl(t,r)Yl(θ),\Phi=\sum_{l=0}^{\infty}\Phi_{l}(t,r)\,Y_{l}(\theta), (91)

where the real spherical harmonics Yl(θ)Y_{l}(\theta) are eigenfunctions of the Laplace-Beltrami operator on S(n1)S^{(n-1)},

Δ̊(n1)Yl(θ)=l(2nl)Yl(θ).\mathring{\Delta}^{(n-1)}\,Y_{l}(\theta)=l(2-n-l)Y_{l}(\theta). (92)

We normalise them such that

Sn1YlYl𝑑S(n1)=A(n2)0πYl(θ)Yl(θ)sinn2θθ=δll,\int_{S^{n-1}}Y_{l}Y_{l^{\prime}}dS^{(n-1)}=A^{(n-2)}\int_{0}^{\pi}Y_{l}(\theta)Y_{l^{\prime}}(\theta)\sin^{n-2}\theta\,\rmd\theta=\delta_{ll^{\prime}}, (93)

where the area A(n2)A^{(n-2)} of Sn2S^{n-2} is given in (57). The mode equation now reads

Φl,tt+Φl,rr+n1rΦl,r+l(2nl)r2Φl=0.-\Phi_{l,tt}+\Phi_{l,rr}+\frac{n-1}{r}\Phi_{l,r}+\frac{l(2-n-l)}{r^{2}}\Phi_{l}=0. (94)

Note that for n=3n=3 we recover (86).

In the following we state explicit solutions for n=5n=5. The first three spherical harmonics are

Y0(θ)=322π,Y1(θ)=1522πcosθ,Y2(θ)=218π(5cos2θ1)Y_{0}(\theta)=\frac{\sqrt{3}}{2\sqrt{2}\pi},\quad Y_{1}(\theta)=\frac{\sqrt{15}}{2\sqrt{2}\pi}\cos\theta,\quad Y_{2}(\theta)=\frac{\sqrt{21}}{8\pi}(5\cos^{2}\theta-1) (95)

and the corresponding radial solutions are

Φ0±(t,r)\displaystyle\Phi_{0}^{\pm}(t,r) =\displaystyle= r2F(r±t)r3F(r±t),\displaystyle r^{-2}F^{\prime}(r\pm t)-r^{-3}F(r\pm t), (96)
Φ1±(t,r)\displaystyle\Phi_{1}^{\pm}(t,r) =\displaystyle= r2F′′(r±t)3r3F(r±t)+3r4F(r±t),\displaystyle r^{-2}F^{\prime\prime}(r\pm t)-3r^{-3}F^{\prime}(r\pm t)+3r^{-4}F(r\pm t), (97)
Φ2±(t,r)\displaystyle\Phi_{2}^{\pm}(t,r) =\displaystyle= r2F′′′(r±t)6r3F′′(r±t)+15r4F(r±t)15r5F(r±t).\displaystyle r^{-2}F^{\prime\prime\prime}(r\pm t)-6r^{-3}F^{\prime\prime}(r\pm t)+15r^{-4}F^{\prime}(r\pm t)-15r^{-5}F(r\pm t). (98)

We observe that now Φl=O(r2)\Phi_{l}^{-}=\Or(r^{-2}) as I+\mathrsfs{I}^{+}\,is approached.

A.3 Transformation to hyperboloidal coordinates

From the conformal coordinates t~,r~\tilde{t},\tilde{r} used in the code, we first compute the physical coordinates

r=2nr~C(1r~2),t=t~+n2C2+r2r=\frac{2n\tilde{r}}{C(1-\tilde{r}^{2})},\quad t=\tilde{t}+\sqrt{\frac{n^{2}}{C^{2}}+r^{2}} (99)

so we can compute Φ(t,r)\Phi(t,r) as constructed above. We then form

Φ~=Ω(1n)/2Φ,\tilde{\Phi}=\Omega^{(1-n)/2}\Phi, (100)

where the conformal factor is given by 16.

The multiplication by a negative power of Ω\Omega in (100) may seem problematic at I+\mathrsfs{I}^{+}\,(r=1r=1), where Ω=0\Omega=0. However, from the explicit solutions given in A.1 and A.2, we see that Φ=O(r(1n)/2)\Phi=\Or(r^{(1-n)/2}), and Ω=r~/r=O(r1)\Omega=\tilde{r}/r=\Or(r^{-1}), so Φ~\tilde{\Phi} in (100) is in fact regular at I+\mathrsfs{I}^{+}\,.

For the initial data we also need to compute the field Π~\tilde{\Pi}. By its definition, we have

Π~=α~1(Φ~,t~β~r~Φ,r~)=α~1Ω(1n)/2[Φ,t~β~r~(Φ,r~+1n2(lnΩ),r~Φ)]\tilde{\Pi}=\tilde{\alpha}^{-1}(\tilde{\Phi}_{,\tilde{t}}-\tilde{\beta}^{\tilde{r}}\Phi_{,\tilde{r}})=\tilde{\alpha}^{-1}\Omega^{(1-n)/2}\left[\Phi_{,\tilde{t}}-\tilde{\beta}^{\tilde{r}}\left(\Phi_{,\tilde{r}}+\frac{1-n}{2}(\ln\Omega)_{,\tilde{r}}\Phi\right)\right] (101)

with Ω\Omega, β~r\tilde{\beta}^{r} and α~\tilde{\alpha} given in (16)–(18). Finally we need to express the derivatives of Φ\Phi in terms of derivatives w.r.t. rr and tt:

Φ,r~=rr~Φ,r+tr~Φ,t,Φ,t~=Φ,t,\Phi_{,\tilde{r}}=\frac{\rmd r}{\rmd\tilde{r}}\Phi_{,r}+\frac{\partial t}{\partial\tilde{r}}\Phi_{,t},\qquad\Phi_{,\tilde{t}}=\Phi_{,t}, (102)

where

rr~=2n(1+r~2)C(1r~2)2,tr~=4nr~C(1r~2)2.\frac{\rmd r}{\rmd\tilde{r}}=\frac{2n(1+\tilde{r}^{2})}{C(1-\tilde{r}^{2})^{2}},\qquad\frac{\partial t}{\partial\tilde{r}}=\frac{4n\tilde{r}}{C(1-\tilde{r}^{2})^{2}}. (103)

The quantities Φ,r\Phi_{,r} and Φ,t\Phi_{,t} can be computed directly from the radial solutions (88)–(90) or (96)–(98).

References

References

  • [1] Szpak N, Bizoń P, Chmaj T and Rostworowski A 2009 Linear and nonlinear tails II: exact decay rates in spherical symmetry J. Hyperbolic Differ. Equ. 6(1) 107–125
  • [2] Strauss W and Vazquez L 1978 Numerical solution of a nonlinear Klein-Gordon equation J. Comput. Phys. 28(2) 271–278
  • [3] Colliander J, Simpson G and Sulem C 2010 Numerical simulations of the energy-supercritical nonlinear Schrödinger equation J. Hyperbolic Differ. Equ. 7(2) 279–296
  • [4] Donninger R and Schlag W 2011 Numerical study of the blowup/global existence dichotomy for the focusing cubic nonlinear Klein–Gordon equation Nonlinearity 24(9) 2547
  • [5] Murphy J and Zhang Y 2020 Numerical simulations for the energy-supercritical nonlinear wave equation Nonlinearity 33 6195–6220
  • [6] Frauendiener J 2004 Conformal infinity Living Rev. Relativity 7(1)
  • [7] Bizoń P and Zenginoğlu A 2009 Universality of global dynamics for the cubic wave equation Nonlinearity 22(10) 2473–2485
  • [8] Zenginoğlu A 2008 A hyperboloidal study of tail decay rates for scalar and Yang–Mills fields Class. Quantum Grav. 25(17) 175013
  • [9] Rinne O and Moncrief V 2013 Hyperboloidal Einstein-matter evolution and tails for scalar and Yang-Mills fields Class. Quantum Grav. 30(9) 095009
  • [10] Rácz I and Tóth G Z 2011 Numerical investigation of the late-time Kerr tails Class. Quantum Grav. 28(19) 195003
  • [11] Csizmadia P, László A and Rácz I 2012 On the use of multipole expansion in time evolution of nonlinear dynamical systems and some surprises related to superradiance Class. Quantum Grav. 30(1) 015010
  • [12] Harris C R et al. 2020 Array programming with NumPy Nature 585(7825) 357–362 https://numpy.org
  • [13] Virtanen P et al. 2020 SciPy 1.0: Fundamental Algorithms for Scientific Computing in Python Nature Methods 17 261–272 https://scipy.org
  • [14] Boyd J P 1978 The Choice of Spectral Functions on a Sphere for Boundary and Eigenvalue Problems: A Comparison of Chebyshev, Fourier and Associated Legendre Expansions Monthly Weather Rev. 106 1184–1191
  • [15] Kreiss H O and Oliger J 1973 Methods for the approximate solution of time dependent problems (Global Atmospheric Research Programme Publication Series no 10) (Geneva: International Council of Scientific Unions, World Meteorological Organization)
  • [16] Boyd J P 2000 Chebyshev and Fourier Spectral Methods (Dover Publications)
  • [17] Fornberg B 1998 A practical guide to pseudospectral methods Cambridge Monographs on Applied and Computational Mathematics (Cambridge University Press)
  • [18] Glogić I, Maliborski M and Schörkhuber B 2020 Threshold for blowup for the supercritical cubic wave equation Nonlinearity 33(5) 2143