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

GWsim: a code to simulate gravitational waves propagating in a potential well

Jian-hua He,1,2
1School of Astronomy and Space Science, Nanjing University, Nanjing 210023, P. R. China
2Key Laboratory of Modern Astronomy and Astrophysics (Nanjing University), Ministry of Education, Nanjing 210023, China

E-mail: [email protected]
(Accepted XXX. Received YYY; in original form ZZZ)
Abstract

We present a code to simulate the propagation of GWs in a potential well in the time domain. Our code uses the finite element method (FEM) based on the publicly available code deal.ii. We test our code using a point source monochromatic spherical wave. We examine not only the waveform observed by a local observer but also the global energy conservation of the waves. We find that our numerical results agree with the analytical predictions very well. Based on our code, we study the propagation of the leading wavefront of GWs in a potential well. We find that our numerical results agree with the results obtained from tracing null geodesics very well. Based on our simulations, we also test the accuracy of the thin-lens model in predicting the positions of the wavefront. We find that the analytical formula of the Shapiro-time delay is only accurate in regimes that are far away from the center of the potential well. However, near the optic axis, the analytical formula shows significant differences from the simulated ones. Besides these results, we find that unlike the conventional images in geometric optics, GWs can not be sheltered by the scatterer due to wave effects. The signals of GWs can circle around the scatterer and travel along the optic axis until arrive at a distant observer, which is an important observational consequence in such a system.

keywords:
gravitational waves – wave effects – wavefront effects
pubyear: 2021pagerange: GWsim: a code to simulate gravitational waves propagating in a potential wellC

1 Introduction

The landmark discovery of gravitational waves Abbott et al. (2016), ripples in the fabric of spacetime triggered by the merger of two black holes, confirms a major prediction of Einstein’s general relativity. It marks the climax of century-long speculation and decades-long painstaking work for hunting such waves. The direct detection of gravitational waves ushered us into a new era of astronomy, allowing us to probe the Universe in an unprecedented way.

In the coming decades, ground- and space-based GW experiments will continue to study GWs, such as the Einstein Telescope (ET) Punturo et al. (2010), 40-km LIGO Dwyer et al. (2015), eLISA eLISA Consortium et al. (2013), DECIGO Sato et al. (2009), and Pulsar Timing Arrays (PTA) Hobbs et al. (2010). The range of the frequency of GWs will be explored from ultra-low frequency regimes such as 109Hz108Hz10^{-9}{\rm Hz}-10^{-8}{\rm Hz} (PTA), 104Hz101Hz10^{-4}{\rm Hz}-10^{-1}{\rm Hz} (eLISA) ,and 102Hz1Hz10^{-2}{\rm Hz}-1{\rm Hz} (DECIGO) to high frequency regimes such as 10Hz104Hz10{\rm Hz}-10^{4}{\rm Hz} (ET and 40-km LIGO). These experiments will usher us into an era of routine GW astronomy and enable us to study the GW phenomenon in unprecedented detail. The combination of these experiments promises to address a variety of outstanding problems in astrophysics, cosmology, and particle physics.

The low-frequency GWs observed by space-borne GW interferometer may shed light on the population of supermassive BHs (SMBHs) with a mass >106M>10^{6}M_{\odot}, which is believed to ubiquity reside in the centers of galaxies Bellovary et al. (2016); Bartos et al. (2017); Stone et al. (2017). SMBHs are major cosmic players and are important to galaxy formation Haehnelt & Kauffmann (2000). The associated jets and accretion disks are the most energetic phenomena in the Universe, which often outshine the entire host galaxy. Therefore, their properties play an important role in the dynamics of stars and gas at galactic scales Ferrarese & Merritt (2000). However, the details of the birth and growth of SMBHs are not well understood yet. The observation of low-frequency GWs, such as in the millihertz band, can probe the spacetime geometry of black hole binaries and their surroundings, which opens a window that can inform us of the properties of these BHs throughout cosmic history. GW observations, thus, can help us to answer many fundamental questions, as to, where and when do the first massive black holes form, how do they grow and assemble, and how their properties are connected to the properties of their host galaxies?

However, in order to achieve these scientific goals, it is crucial to accurately infer the physical properties of the distant source from the observed GW signals. This process, however, is likely beset by the wave nature of GWs. Unlike the visible light, the wavelength of the low-frequency GWs is much larger. In the millihertz band, for example, the wavelength can be as large as several times the astronomical unit (AU{\rm AU}). When the wavelength of the GWs is comparable to the Schwarzschild radius of an object, its propagation no longer obeys the geometrical optics. Further, if the gravitational waves come from the same compact binaries, they should be coherent and interference. In all these cases, the wave effects of GWs are significant and have to be taken into account Bontz & Haugan (1981); Ohanian (1974).

The wave effects in a lensing system have already been studied in the previous work Deguchi & Watson (1986a, b); Schneider et al. (1992); Ruffa (1999); De Paolis et al. (2002); Takahashi & Nakamura (2003); Nakamura & Deguchi (1999); Suyama et al. (2005); Christian et al. (2018); Zakharov & Baryshev (2002); Liao et al. (2019); Macquart (2004); Nakamura (1998); Dai et al. (2018); Cao et al. (2014); Yoo et al. (2013); Nambu et al. (2019). In these pioneer work (e.g. Schneider et al. (1992)), the lensing system is usually considered as a thin lens model: the gravitational field of the compact lens is weak, and the deflection angle due to the lens masses is small. The impact parameter of the incident rays is thought to be much larger than the Schwarzschild radius of the lens mass. As such, by construction the thin-lens model does not work for light rays that are close to the optic axis. Moreover, the changes in the amplitude of the incident waves are negligible at the location of the lens mass. The changes in the amplitude, however, are mainly due to the deflection near the lens, which is small at the beginning but becomes significant after the long journey of the light rays from the lens to the observer (see discussions in (4.86) in  Schneider et al. (1992)).

In addition to the thin-lens model, the previous work also applied Kirchhoff’s diffraction theory to the gravitational lensing system Schneider et al. (1992). The lens is considered as an optic aperture. The lensed signal is then thought to be the diffraction of the waves from the aperture. An important assumption in Kirchhoff’s theory is that boundaries outside the aperture are set to be zero. Kirchhoff’s theory therefore assumes that only waves re-radiated from the aperture can be observed by the observer. The diffraction integral over the aperture, indeed, only considers the scattered waves, which neglects the contribution of the original incident waves (see Eq.(5) in  Takahashi et al. (2005)). This is valid in the system of wave optics, as in this case the original incident waves are blocked by obstacles outside the aperture. However, if the amplitude of the scattered wave is comparable to that of the incident wave, the original incident waves have to be considered as well (This similar to the Babinet’s principle in wave optics, in which the aperture is replaced by a screen. In this case, the incident and scattered waves have to be considered simultaneously.)

If we consider the incident waves, the lensing system can be treated as a scattering system Weinberg (1972); Peters (1974); Takahashi et al. (2005); Sorge (2015). Indeed, the scattering of waves by potential wells has been well studied in Quantum Mechanics. As shown in  Peters (1974), the wave equation of GWs propagating in a non-uniform background spacetime is similar to that of the well-known Schrödinger’s equation for the scattering of two charged particles. Moreover,  Peters (1974); Takahashi et al. (2005) also generalized the scattering model that uses a plane wave as the incident wave to a spherical one.

Nevertheless, despite these progresses, there are still some concerns. The previous analyses mainly focus on the frequency-domain with implicit assumptions, such as that GWs are in the steady-state (e.g. Peters (1974)) or stationary state (e.g. Suyama et al. (2005)). These analyses also assume that the wave zone of GWs is infinite, which neglects the causality and wavefront effects. However, the realistic GW signals are time-finite and, indeed, neither in a steady-state nor stationary state. Moreover, unlike in the 2D case (or even dimensions), a prominent feature of 3D (or odd dimensions) waves is that they follow the strong Huygens’ principle. In free space, a perturbation in 3D space at a point x\vec{x} is visible at another point x\vec{x}^{\prime} exactly at the time t=|xx|/ct=|\vec{x}-\vec{x}^{\prime}|/c but not later. As a result, a time-finite wave signal has a definite wave zone with clear leading and trailing wavefronts as boundaries. As the leading wavefront represents the transport of energy from one place to another, it has to follow the constraint of causality. The behavior of the leading wavefront, therefore, is different from the rest part of the waves. Since GWs have a finite wave zone, the wavefront effects should be taken into account. However, despite their importance, they have not yet been properly addressed in the literature.

On the other hand, to effectively address these issues is, indeed, highly non-trivial. First, the leading wavefront is not necessarily to be continuous but can be a step-like function with a sharp edge (e.g. square waves). In this case, the sums of the Fourier series do not converge near the point of discontinuity of a piecewise smooth function. Even with a large number of the Fourier series, the sums oscillate at the discontinuity point, which is approximately 9%9\% of the jump of the discontinuity111In mathematics, the Fourier series converges only in the sense of L2L_{2} norm, which means that they converge up to functions with differences on a set of Lebesgue measure zero. However, they do not necessarily converge in the pointwise sense, unless the function is continuously differentiable.. This phenomenon is known as the Gibbs phenomenon. Second, the wavefront may have a complex shape in a 3D potential well, which can lead to complex boundaries of the wave zones. The domain of the wave zone might even not be convex222The solutions of partial differential equations (PDEs) are strongly dependent on the shape of the boundaries of the domain. It is offten assumed that the shape of boundaries is regular or of Lipschitz continuity. As such, the domains of PDEs are usually assumed to be bounded convex domains.. However, even for the simplest Laplacian operator, if the dimension of space is greater than 1, the eigenvalues and eigenfunctions are known only for simple geometries (see detailed discussions in chapter 1 of Grossmann et al. (2007)). As such, in general cases, it is difficult to analyze the behaviors of the wavefront in a potential well using the spectrum method based on eigenfunctions (e.g. Fourier analysis). Therefore, numerical techniques are called for in such situations.

This work aims to develop a code that can efficiently simulate the propagation of GWs in a potential well. The main technique used in this work is called the finite element method (FEM). The FEM is an effective and well-developed method for numerically solving partial differential equations (PDEs) (see e.g. text book Zienkiewicz & Taylor (1989) for details). Unlike most of the conventional methods, such as the finite difference method (FDM), the FEM is based on the weak formulation of PDEs. The basic idea of the FEM is to present the domain of interest as an assembly of finite elements. On each finite element, the solution of PDEs is approximated by local shape functions. A continuous PDE problem can then be transformed into a discretized finite element with unknown nodal values. These unknowns form a system of linear algebraic equations, which can be solved numerically. Compared to the conventional numerical methods, such as the FDM, which requires regular meshes, the advantage of the FEM is that it can work for complex boundaries so that it can significantly minimize the effects of boundaries. Moreover, it can also provide very good precision even with simple approximation shape functions. Due to the locality of the approximation shape functions, the resulting linear algebraic equations are sparse, which can be solved efficiently in the numerical process.

Throughout this paper, we adopt the geometric unit c=G=1c=G=1, in which 1Mpc=1.02938×1014Sec1\,{\rm Mpc}=1.02938\times 10^{14}{\rm Sec} and 1M=4.92535×106Sec.1M_{\odot}=4.92535\times 10^{-6}{\rm Sec}\,. The advantage of this unit system is that time and space have the same unit.

This paper is organized as follows: In Section 2, we introduce the Strong and Weak formulation of the wave equation. In Section 3, we describe the FEM for wave equations. In Section 4, we present several tests of our code using spherical waves. In Section 5, we simulate GWs in a potential well. In Section 6, we summarize and conclude this work.

2 Strong and Weak formulation of the wave equation

In this section, we present an overview of the mathematical basis of the wave equation. To begin with, we discuss the strong formulation of the wave equation and then we will introduce the weak formulation.

2.1 Strong formulation

Let Ω3\Omega\subset\mathbb{R}^{3} be a bounded domain with boundary Ω\partial\Omega. The boundary value problem (BVP) we aim to study in this work can be presented as

c22u2t2u=c2finΩ×(0,T],c^{2}\nabla^{2}u-\frac{\partial^{2}}{\partial t^{2}}u=c^{2}f\quad{\rm in}\quad\Omega\times(0,T]\,, (1)

which is a second-order linear hyperbolic partial differential equation (PDE) with initial conditions

u(x,0)=u0(x)inΩ.\displaystyle u(x,0)=u_{0}(x)\quad{\rm in}\quad\Omega\,. (2)

Let Ω1\partial\Omega_{1} and Ω2\partial\Omega_{2} are subsets of the boundary Ω\partial\Omega with Ω1Ω2=Ω\partial\Omega_{1}\cup\partial\Omega_{2}=\partial\Omega and Ω1Ω2=\partial\Omega_{1}\cap\partial\Omega_{2}=\emptyset. We assume that the boundary conditions on Ω1\partial\Omega_{1} are Dirichlet and on Ω2\partial\Omega_{2} is Neumann

u(x,t)\displaystyle u(x,t) =u1(x,t)\displaystyle=u_{1}(x,t) on\displaystyle{\rm on} Ω1×(0,T]\displaystyle\quad\partial\Omega_{1}\times(0,T] (3)
n^u\displaystyle\hat{n}\cdot\nabla u =q(x,t)\displaystyle=q(x,t) on\displaystyle{\rm on} Ω2×(0,T],\displaystyle\quad\partial\Omega_{2}\times(0,T]\,, (4)

where n^\hat{n} is the outward pointing unit vector normal to Ω2\partial\Omega_{2}. q(x,t)q(x,t) is a given function on the boundary Ω2\partial\Omega_{2}.

In mathematics, Eq. (1), together with the initial and boundary conditions Eqs. (2,3,4), are called the strong formulation. The strong formulation is the most commonly used format of the wave equation. It can be numerically solved using some straightforward methods such as the finite difference method (FDM). However, the strong formulation is not the only format of the wave equation. In what follows, we introduce an alternative format, namely, the weak formulation of the above equations.

2.2 Weak formulation

Multiply Eq. (1) by a test function ϕ\phi and then integrate over Ω\Omega, we obtain a variational equation

Ωu(c2ϕ)dx+Ω(n^u)(c2ϕ)dx2t2Ωuϕdx=Ωc2fϕdxϕV,\displaystyle\begin{split}&-\int_{\Omega}\nabla u\cdot\nabla(c^{2}\phi)\,\mathrm{d}x+\int_{\partial\Omega}(\hat{n}\cdot\nabla u)(c^{2}\phi)\,\mathrm{d}x-\frac{\partial^{2}}{\partial t^{2}}\int_{\Omega}u\phi\,\mathrm{d}x\\ =&\int_{\Omega}c^{2}f\phi\,\mathrm{d}x\quad\forall\phi\in V\end{split}\,, (5)

where we have used the Green’s formula

Ωc2ϕ2udx=Ω(c2ϕ)udx+Ωc2ϕn^udx.\int_{\Omega}c^{2}\phi\nabla^{2}u\,\mathrm{d}x=-\int_{\Omega}\nabla(c^{2}\phi)\cdot\nabla u\,\mathrm{d}x+\int_{\partial\Omega}c^{2}\phi\hat{n}\cdot\nabla u\,\mathrm{d}x\,.

The second term on the right hand represents the boundary term, which can be further split into terms according to different boundary conditions

Ω(n^u)(c2ϕ)dx=Ω1(n^u)(c2ϕ)dx+Ω2(n^u)(c2ϕ)dx.\displaystyle\begin{split}\int_{\partial\Omega}(\hat{n}\cdot\nabla u)(c^{2}\phi)\,\mathrm{d}x&=\int_{\partial\Omega_{1}}(\hat{n}\cdot\nabla u)(c^{2}\phi)\,\mathrm{d}x\\ &+\int_{\partial\Omega_{2}}(\hat{n}\cdot\nabla u)(c^{2}\phi)\,\mathrm{d}x\end{split}\,. (6)

The test function ϕ\phi is chosen in such a way that it vanishes on the subset of boundaries with Dirichlet boundary conditions Ω1{\partial\Omega_{1}}, which is equivalent to say that we restrict ϕ\phi to lie in

V:={ϕ:ϕH1(Ω),ϕ|Ω1=0},V:=\{\phi:\phi\in H^{1}(\Omega),\phi|_{\partial\Omega_{1}}=0\}\,, (7)

where H1(Ω)=W1,2(Ω)H^{1}(\Omega)=W^{1,2}(\Omega) is called the first order Sobolev space, which is also a Hilbert space, meaning that ϕ\phi and its first order weak derivatives xϕ\partial_{x}\phi are square integrable.

ϕH1(Ω)=[Ω|α|1|xαϕ(x)|2dx]12<.\left\|\phi\right\|_{H^{1}(\Omega)}=\left[\int_{\Omega}\sum_{|\alpha|\leq 1}|\partial^{\alpha}_{x}\phi(x)|^{2}dx\right]^{\frac{1}{2}}<\infty\,.

Given the restrictions of the test function ϕ\phi, the boundary term in Eq. (5) can be reduced to

Ω(n^u)(c2ϕ)dx=Ω2(n^u)(c2ϕ)=Ω2ut(cϕ)dx.\displaystyle\begin{split}\int_{\partial\Omega}(\hat{n}\cdot\nabla u)(c^{2}\phi)\,\mathrm{d}x&=\int_{\partial\Omega_{2}}(\hat{n}\cdot\nabla u)(c^{2}\phi)\\ &=-\int_{\partial\Omega_{2}}\frac{\partial{u}}{\partial{t}}(c\phi)\,\mathrm{d}x\end{split}\,. (8)

Here, in the second equality, we have imposed an absorbing boundary condition

n^u=1cutonΩ2×(0,T].\hat{n}\cdot\nabla u=-\frac{1}{c}\frac{\partial{u}}{\partial{t}}\quad{\rm on}\quad\partial\Omega_{2}\times(0,T]\,. (9)

The physical meaning of the absorbing boundary condition will be explained later on.

The above we have discussed the boundary condition on Ω2\partial\Omega_{2}. On the other hand, the boundary condition on Ω1\partial\Omega_{1},

{u(x,t)=u0(x,t)u(x,t)t=u0(x,t)tonΩ1×(0,T],\left\{\begin{aligned} u(x,t)&=u_{0}(x,t)\\ \frac{\partial u(x,t)}{\partial t}&=\frac{\partial u_{0}(x,t)}{\partial t}\end{aligned}\right.\quad{\rm on}\quad\partial\Omega_{1}\times(0,T]\,, (10)

is called the essential boundary condition, which does not affect the variational equation Eq. (5) but this condition must be imposed directly on the solution itself.

Equipped with the boundary conditions Eqs. (9,10), Eq. (5) is called the weak formulation of Eq. (1). It can be shown that if Eq. (5) holds for all ϕV\phi\in V, it has a unique solution uVu\in V Lax & Milgram (1954). This solution is called the weak solution. It is obvious that the classical solutions of Eq. (1) are also weak solutions. However, conversely, weak solutions must have sufficient smoothness to be the classical solutions, which dependents both on the shape of the boundary Ω\partial\Omega and the regularity of c2fc^{2}f. For example, it can be shown that if Ωn\Omega\subset\mathbb{R}^{n} has CkC^{k} boundary (smooth up to kk-th derivative) and c2fHk(Ω)c^{2}f\in H^{k}(\Omega) with k>n2k>\frac{n}{2}, then uC2(Ω¯)u\in C^{2}(\bar{\Omega}), namely uu is a solution in the classical sense Grisvard (1985).

2.3 Energy Conservation of the wave equation

In addition to the weak formulation of the wave equation, another important property of the wave function Eq. (1) is that it obeys the law of energy conservation. To see this point, we multiply Eq. (1) by ut\frac{\partial u}{\partial t} and integrate over the domain Ω\Omega

12tΩ[(ut)2+c2(u)2]dx=Ωc2(ut)(n^u)dxΩ(ut)[(c2)u+c2f]dx.\displaystyle\begin{split}&\frac{1}{2}\frac{\partial}{\partial t}\int_{\Omega}\left[\left(\frac{\partial{u}}{\partial t}\right)^{2}+c^{2}\left(\nabla u\right)^{2}\right]\,\mathrm{d}x\\ =&\int_{\partial\Omega}c^{2}\left(\frac{\partial u}{\partial t}\right)(\hat{n}\cdot\nabla u)\,\mathrm{d}x-\int_{\Omega}\left(\frac{\partial u}{\partial t}\right)\left[\left(\nabla c^{2}\right)\cdot\nabla u+c^{2}f\right]\,\mathrm{d}x\end{split}\,. (11)

Note that the total energy carried by a wave in the domain Ω\Omega is defined by

E(t)=12Ω[(ut)2+c2(u)2]dx.E(t)=\frac{1}{2}\int_{\Omega}\left[\left(\frac{\partial{u}}{\partial t}\right)^{2}+c^{2}\left(\nabla u\right)^{2}\right]\,\mathrm{d}x\,. (12)

The left hand of the equality of Eq. (11) then represents the change of the total energy of wave with time. From Eq. (11), if the boundary condition on the right hand side of the equality is a constant (ut|Ω=0\frac{\partial{u}}{\partial t}|_{\partial\Omega}=0) and there is no force source f=0f=0 in the second term, for a constant speed of wave c2=0\nabla c^{2}=0, the total energy is conserved dE(t)dt=0\frac{dE(t)}{dt}=0.

Equation (11) is an important feature of the wave function. We shall discuss this point in our following numerical analyses.

3 The Finite Element Method

Before presenting our numerical analyses, we introduce an auxiliary function v:=utv:=\frac{\partial u}{\partial t} and adopt the common notation

(f,g)Ω=Ωf(x)g(x)dx,\left(f,g\right)_{\Omega}=\int_{\Omega}f(x)g(x)\,\mathrm{d}x\quad, (13)

for convenience. Equation (5) in this case can be re-written as a set of two equations

(ut,ϕ)Ω(v,ϕ)Ω=0(u,(c2ϕ))Ω+(ut,cϕ)Ω+(vt,ϕ)Ω+(c2f,ϕ)Ω=0}ϕV.\left.\begin{aligned} \left(\frac{\partial u}{\partial t},\phi\right)_{\Omega}-\left(v,\phi\right)_{\Omega}&=0\\ \left(\nabla u,\nabla(c^{2}\phi)\right)_{\Omega}+\left(\frac{\partial u}{\partial t},c\phi\right)_{\partial\Omega}+\left(\frac{\partial v}{\partial t},\phi\right)_{\Omega}+\left(c^{2}f,\phi\right)_{\Omega}&=0\end{aligned}\right\}\quad\forall\phi\in V\,. (14)

To solve the above equations numerically, we need to discretize these equations first. In general for a time-dependent problem, the discretization can be done either for the spatial variables first or the temporal variables first. The method to discretize the spatial variables first is called the (vertical) method of lines (MOL). Once the spatial variables are discretized, each variable needs to be solved evolving with time. This constitutes a large system of ordinary differential equations (ODEs). The advantage of this method is that ODEs can be numerically solved by well-developed ODE solvers that are efficient and of high-order accuracy(see discussions in Chapter 5 of Ref. Grossmann et al. (2007)).

The alternative approach is called the Rothe’s method, in which the temporal variable is discretized first. The advantage of this method is that the spatial resolution can be changed at each time step, which allows for the usage of the technique of adaptive mesh refinement (AMR). However, when trying to apply high-order temporal discretization schemes, Rothe’s method is rather awkward.

Whether the time variable or the spatial variable should be discretized first has long been debated in the numerical analysis community. In this work, however, we follow the Rothe’s method as we shall show that the spatial resolution is, indeed, the major limiting factor for the problem that we are interested in. The Rothe’s method allows for great freedom to deal with the spatial resolution.

3.1 discretization in time

We use superscript nn to indicate the number of a time step and k=tntn1k=t_{n}-t_{n-1} is the length of the present time step. We discretize the time derivative as

{vtvnvn1kutunun1k.\displaystyle\left\{\begin{aligned} \frac{\partial v}{\partial t}&\approx\frac{v^{n}-v^{n-1}}{k}\quad\\ \frac{\partial u}{\partial t}&\approx\frac{u^{n}-u^{n-1}}{k}\end{aligned}\right.\,. (15)

The discretization scheme we adopted here also involves the spatial quantities at two different time steps, which is known as the θ\theta-scheme. Equations (14) then can be represented as

(unun1k,ϕ)Ω(θvn+(1θ)vn1,ϕ)Ω=0;([θun+(1θ)un1],(c2ϕ))Ω(unun1k,cϕ)Ω=(vnvn1k,ϕ)Ω+(c2[θfn+(1θ)fn1],ϕ)Ω}ϕV.\left.\begin{aligned} &\left(\frac{u^{n}-u^{n-1}}{k},\phi\right)_{\Omega}-\left(\theta v^{n}+(1-\theta)v^{n-1},\phi\right)_{\Omega}=0\,;\\ &-\left(\nabla[\theta u^{n}+(1-\theta)u^{n-1}],\nabla(c^{2}\phi)\right)_{\Omega}-\left(\frac{u^{n}-u^{n-1}}{k},c\phi\right)_{\partial\Omega}\\ =&\left(\frac{v^{n}-v^{n-1}}{k},\phi\right)_{\Omega}+\left(c^{2}[\theta f^{n}+(1-\theta)f^{n-1}],\phi\right)_{\Omega}\end{aligned}\right\}\quad\forall\phi\in V\,. (16)

When θ=0\theta=0, the scheme is called the forward or explicit Euler method. If θ=1\theta=1, it reduces to the backward or implicit Euler method. The Euler method, no matter implicit or explicit, is only of first-order accurate. Therefore we do not adopt here.

The scheme we use here is called the Crank-Nicolson Scheme, namely θ=1/2\theta=1/2, which uses the midpoint between two different time steps. This scheme is of second-order accuracy. As the midpoint rule is symplectic, the most important feature of this scheme is that it is energy-preserving. We shall discuss this point in our following numerical analyses.

3.2 discretization in space

We discretize the domain Ω3\Omega\subset\mathbb{R}^{3} using the Finite Elements Methods (FEM). The FEM is an effective method for numerically solving partial differential equations. There are many excellent textbooks on this topic (Readers are referred to e.g. Ref. Zienkiewicz & Taylor (1989) for more details). Here, we will not go into too many mathematical details. Instead, we only summarise the main steps here.

First, the domain Ω\Omega is decomposed into subdomains Ωi\Omega_{i}, which consist of rectangles and triangles. This is called decomposition or triangulation. The vertices of rectangles and triangles in the domain Ω\Omega are called mesh points or nodes. Let Ωh\Omega_{h} denote the set of all nodes of the decomposition. On each node, we construct a test function ϕiV,i=1,..,N\phi_{i}\in V\,,i=1,..,N, where VV is the space defined in Eq. (7) and NN is the total number of nodes in the domain. The test function ϕi\phi_{i} is required to have the property

ϕi(pk)=δik,i,k=1,..,N,pkΩh,\phi_{i}(p^{k})=\delta_{ik},\quad i,k=1,..,N,\quad p^{k}\in\Omega_{h}\,,

where

δik={1fori=k0forik.\delta_{ik}=\left\{\begin{aligned} 1\quad&{\rm for}&\quad i=k\\ 0\quad&{\rm for}&\quad i\neq k\end{aligned}\right.\quad.

Thus, ϕi\phi_{i} has non-zero values only on the node with k=ik=i and its adjacent subdomains. It vanishes on other parts of the domain Ω\Omega. The test function ϕi\phi_{i} constructed this way is called the shape function. Clearly, ϕiV\phi_{i}\in V on different nodes are linearly independent. We denote the space spanned by ϕi\phi_{i} as Vh:=span{ϕi}i=1NV_{h}:=\mathrm{span}\{\phi_{i}\}_{i=1}^{N}, which is a subspace of VV.

Next, we expand the scalar fields u,v,c2fu,v,c^{2}f in terms of ϕi\phi_{i}

uh=i=1Nuiϕi,vh=i=1Nviϕi,c2fh=i=1Nc2fiϕi,u_{h}=\sum_{i=1}^{N}u_{i}\phi_{i}\,,v_{h}=\sum_{i=1}^{N}v_{i}\phi_{i}\,,c^{2}f_{h}=\sum_{i=1}^{N}c^{2}f_{i}\phi_{i}\,, (17)

where uh,vh,c2fhVhu_{h}\,,v_{h}\,,c^{2}f_{h}\in V_{h}. These functions are approximations of the original scalar fields. The advantage of this expansion is that the coefficients are precisely the values of these functions on the node points (e.g. ui=uh(pi)u_{i}=u_{h}(p_{i})).

Inserting the above expansions into Eqs. (16) and noting that Eqs. (16) holds for any ϕV\phi\in V, thus, we can choose ϕ\phi as ϕ=ϕj\phi=\phi_{j}, where j=1,..,Nj=1,..,N. Then we obtain NN different equations, which form a linear system

[kθ(A+D)+B]Un\displaystyle[k\theta(A+D)+B]U^{n} =MVn+G2\displaystyle=-MV^{n}+G_{2}
k[θFn+(1θ)Fn1],\displaystyle-k[\theta F^{n}+(1-\theta)F^{n-1}]\,, (18)
MUn\displaystyle MU^{n} =kθMVn+G1,\displaystyle=k\theta MV^{n}+G_{1}\,, (19)

where G1G_{1} and G2G_{2} are defined by

{G1=MUn1+k(1θ)MVn1G2=[k(1θ)(A+D)+B]Un1+MVn1,\left\{\begin{aligned} G_{1}&=MU^{n-1}+k(1-\theta)MV^{n-1}\\ G_{2}&=[-k(1-\theta)(A+D)+B]U^{n-1}+MV^{n-1}\end{aligned}\right.\,,

and the elements of the matrixes are defined by

{Dij=((c2)ϕi,ϕj)ΩAij=(c2ϕi,ϕj)ΩMij=(ϕi,ϕj)ΩBij=(cϕi,ϕj)ΩFin=(c2fn,ϕi)ΩUin=(un,ϕi)ΩVin=(vn,ϕi)Ω.\left\{\begin{aligned} D_{ij}&=\left(\nabla(c^{2})\phi_{i},\nabla\phi_{j}\right)_{\Omega}\\ A_{ij}&=\left(c^{2}\nabla\phi_{i},\nabla\phi_{j}\right)_{\Omega}\\ M_{ij}&=\left(\phi_{i},\phi_{j}\right)_{\Omega}\\ B_{ij}&=\left(c\phi_{i},\phi_{j}\right)_{\partial\Omega}\\ F^{n}_{i}&=\left(c^{2}f^{n},\phi_{i}\right)_{\Omega}\\ U^{n}_{i}&=\left(u^{n},\phi_{i}\right)_{\Omega}\\ V^{n}_{i}&=\left(v^{n},\phi_{i}\right)_{\Omega}\end{aligned}\right.\,.

It is worth noting that the wave speed cc is not a constant so that the matrix DD is not symmetric DijDjiD_{ij}\neq D_{ji}, which needs to be kept consistent with Eqs. (17). Equations (18,19) are called the Galerkin equations, from which the unknown coefficients uiu_{i} and viv_{i} can be solved. Note that Eq. (18) explicitly contains VnV^{n}, which is unknown at the time step nn. However, we can multiply Eq. (18) by kθk\theta and then use Eq. (19) to eliminate VnV^{n}. Then we obtain

[M+k2θ2(A+D)+kθB]Un\displaystyle[M+k^{2}\theta^{2}(A+D)+k\theta B]U^{n} =G1+kθG2\displaystyle=G_{1}+k\theta G_{2}
k2θ[θFn+(1θ)Fn1],\displaystyle-k^{2}\theta[\theta F^{n}+(1-\theta)F^{n-1}]\,, (20)

for UnU^{n} and

MVn\displaystyle MV^{n} =[kθ(A+D)+B]Un+G2\displaystyle=-[k\theta(A+D)+B]U^{n}+G_{2}
k[θFn+(1θ)Fn1],\displaystyle-k[\theta F^{n}+(1-\theta)F^{n-1}]\,, (21)

for VnV^{n}.

Now, UnU^{n} can be solved using the information on the time step n1n-1. Then VnV^{n} can be solved with the knowledge of UnU^{n}. In practice, Eqs. (20,21) usually contain a very large number of equations. The rank of Eqs. (20,21) (the number of linear independent equations) is called the degrees-of-freedom (DOF) of the system, which depends both on the number of node points and the freedoms in the shape functions.

3.3 linear solvers

As the systems of linear Eqs. (20,21) may contain a large number of unknowns(e.g. which can be easily up to a few 10710^{7}), direct methods such as the LU decomposition is inefficient in this case. One has to use iterative methods such as the Conjugate Gradient (CG) method, which is efficient for large but sparse matrices. However, our cases are complicated due to the varying wave speed cc. The matrix DD is no longer symmetric. As the CG method can be only applied to symmetric, positive-definite matrices, we can not use the CG method in this work. Instead, we adopt the GMRES (a generalized minimal residual algorithm for solving non-symmetric linear systems) method Young & Chen (1996), which does not require any specific properties of the matrices.

3.4 numerical implementation

In this work, we use the public available code deal.ii Arndt et al. (2019); Bangerth et al. (2007); Alzetta et al. (2018) to numerically solve Eqs. (20,21). deal.ii is a C++ program library aimed at numerically solving partial differential equations using modern finite element method. Coupled to stand-alone linear algebra libraries, such as PETSc Abhyankar et al. (2018); Balay et al. (2019a, b, 1997), deal.ii supports massively parallel computing of vary large linear systems of equations. deal.ii also provides convenient tools for triangulation of various geometries of the simulation domain. Therefore, it is straightforward to implement Eqs. (20,21) in the deal.ii code. The detailed numerical analyses are presented in the next section.

4 Code Tests

In this section, we present several tests of our code using an important physical model that describes waves emitting from a point source. Moreover, we shall discuss the Huygens-Fresnel principle for wave functions. We use the Huygens-Fresnel principle to set boundary conditions so that we can avoid mathematical singularities of spherical waves at the origin.

4.1 Spherical Waves from a Point Source and the Huygens-Fresnel principle

The boundary-initial value problem for the propagation of waves from a point source can be presented as

2u1c22t2u=Q(t)δ(x)inΩ×(0,T],\nabla^{2}u-\frac{1}{c^{2}}\frac{\partial^{2}}{\partial t^{2}}u=-Q(t)\delta(x)\quad{\rm in}\quad\Omega\times(0,T]\,, (22)

where Q(t)Q(t) is the time-dependent waveform of the point source and δ\delta is the Dirac delta function. The wave equation has a fundamental solution, which is the distributional solution of

2G1c22t2G=δ(t)δ(x).\nabla^{2}G-\frac{1}{c^{2}}\frac{\partial^{2}}{\partial t^{2}}G=-\delta(t)\delta(x)\,. (23)

The distribution GG is also called the Green’s function. In free-space, the explicit form of GG for an outgoing wave is given by

G+(x,t;x,t)=δ(t[txxc])4πxx.G^{+}(x,t;x^{\prime},t^{\prime})=\frac{\delta(t^{\prime}-[t-\frac{\left\|x-x^{\prime}\right\|}{c}])}{4\pi\left\|x-x^{\prime}\right\|}\,. (24)

The above equation is known as the retarded Green’s function in free-space. The general solution of Eq. (22) then has the form

u(x,t)=G+(x,t;x,t)Q(t)δ(x)𝑑x𝑑t=Q(trc)4πr,\displaystyle\begin{split}u(x,t)&=\int\int G^{+}(x,t;x^{\prime},t^{\prime})Q(t^{\prime})\delta(x^{\prime})dx^{\prime}dt^{\prime}\\ &=\frac{Q(t-\frac{r}{c})}{4\pi r}\,,\end{split} (25)

where rr is the norm of xx, namely, r=xr=\left\|x\right\|.

Although the wave function from a point source has a very simple analytic expression Eq. (25), numerically solving Eq. (22) is, indeed, non-trivial. A challenge is that the point source on the right-hand side of Eq. (22) is singular, which may cause numerical problems. In practice, in order to avoid this problem, we can make use of the Huygens-Fresnel principle to numerically solve Eq. (22) on a slightly different domain without the origin but with new boundaries

2u1c22t2u\displaystyle\nabla^{2}u-\frac{1}{c^{2}}\frac{\partial^{2}}{\partial t^{2}}u =0inΩ/{0}×(0,T],\displaystyle=0\quad{\rm in}\quad\Omega/\{0\}\times(0,T]\,, (26)

The Huygens-Fresnel principle states that the original waves from a source propagating to a distant observer can be considered as re-radiating from a wavefront of the original wave rather than directly from the source. As such, we can use the wavefront of the original wave at some radius as a new boundary for the wave equation. In this way, we can avoid the singularity in Eq. (22) at the origin. A mathematical description of the Huygens-Fresnel principle for spherical waves is provided in Appendix A.

Refer to caption
Figure 1: The boundaries and triangulation of our simulation domain. The domain consists of the bulk regions between the two spherical shells. The plot shows a slice taken through the center of the sphere. The meshes have a resolution of 252^{5}. On Ω1\partial\Omega_{1}, we set the boundary conditions as the wavefront of the original wave from the point source propagating there. However, we set the absorbing boundary condition on Ω2\partial\Omega_{2}.

Figure 1 shows the new boundaries and triangulation of our simulation domain. The domain consists of the bulk regions between the two spherical shells with an inner radius RiR_{i} and an outer radius RoR_{o}, respectively. We set the boundary conditions at the inner shell RiR_{i} as

u(Ri,t)={Q(tRic)4πRitRi/c0t<Ri/conΩ1×(0,T],u(R_{i},t)=\left\{\begin{aligned} &\frac{Q(t-\frac{R_{i}}{c})}{4\pi R_{i}}\quad&t\geq R_{i}/c\\ &0\quad&t<R_{i}/c\end{aligned}\right.\quad{\rm on}\quad\partial\Omega_{1}\times(0,T]\,,\\

which is exactly the wavefront of the original point source propagating there. At the outer shell, we adopt the absorbing boundary condition

ur(Ro,t)=1cutonΩ2×(0,T].\frac{\partial u}{\partial r}(R_{o},t)=-\frac{1}{c}\frac{\partial u}{\partial t}\quad{\rm on}\quad\partial\Omega_{2}\times(0,T]\,. (27)

The absorbing boundary condition is also called the non-reflecting boundary conditions or radiating boundary conditions. These boundary conditions can absorb and eliminate the reflections of waves on the boundaries. Therefore, with these boundary conditions, we can approximate the propagation of waves in free-space using a limited volume of the simulation domain. In addition to the boundary conditions, we set the initial conditions as

u(x,0)=0inΩ.u(x,0)=0\quad{\rm in}\quad\Omega\,. (28)

After fixing the boundary and initial conditions, we perform several numerical tests. In these tests, we choose a concrete waveform Q(t)=Asin(ωt)Q(t)=-A\sin(\omega t). Thus, the spherical wave has an analytical expression

u(x,t)=Asin[ω(tr/c)]4πr,u(x,t)=-\frac{A\sin[\omega(t-r/c)]}{4\pi r}\,, (29)

where ω=2πf\omega=2\pi f\, and ff is the frequency of waves. AA is the amplitude of the wave. In our test, we set Ro=107M=49.2535Sec,R_{o}=10^{7}M_{\odot}=49.2535{\rm Sec}\,,Ri=9×106M=44.32815SecR_{i}=9\times 10^{6}M_{\odot}=44.32815{\rm Sec} and f=1f=1. Further, we set A=4πRiA=4\pi R_{i}, namely, the amplitude of the waves are normalized to unity at the inner boundary.

4.2 The choice of time step

To perform our simulations, we further need to set the time step. For a time-dependent wave equation, an explicit discretization method is only stable if the time step is small enough so that the wave can have enough time to travel through the space discretization (see. Chapter 2 in Ref. Grossmann et al. (2007)). This condition is known as the Courant-Friedrichs-Levi condition

kσc,k\leq\frac{\sigma}{c}\,, (30)

where σ\sigma is the size of the mesh and kk is the size of the time step as mentioned before. However, for the implicit methods such as the one used in this work, it can be stable for arbitrary step sizes if the scheme is upwind (see. Chapter 2 in Ref. Grossmann et al. (2007)). However, a stable scheme is not sufficient to guarantee a correct solution to the wave equation. This is because of the oscillating features of waves. If the frequency of waves is too high, the wavelength will be too small and it is difficult to resolve the waveform within one period. The shape of the waves, in this case, is not smooth relative to the size of the meshes, which can lead to large errors in the numerical simulations. Therefore, in order to get a correct solution, the time step and the size of the meshes should be small enough so that the waveform within one period can be well resolved. As such, in our simulations, we set k=λ/30k=\lambda/30 and the size of the mesh σ<λ/7\sigma<\lambda/7. We have tested that further reducing the size of the mesh σ\sigma does not improve the results appreciably.

Refer to caption
Figure 2: Numerical results (dashed lines) against analytical results (solid lines). The waveform is observed by an observer located at r=46.63Secr=46.63{\rm Sec}. The left vertical dashed line indicates the epoch that the wavefront just arrives at the observer. The right vertical dashed line indicates the epoch that the wavefront arrives at the outer boundary of our simulation domain. After this epoch, the wave starts getting out of the simulation domain. Dashed lines with different colors show the results obtained using different spatial resolutions.

4.3 Numerical results

We perform a suite of simulations with different spatial resolutions. In these simulations, we adopt the linear Lagrange test function, which depends only on the values of the finite elements on their vertices but does not involve any derivatives of the unknown fields. Therefore the degree of freedom (DOF) associated with each vertex is one. In this case, the total DOF is simply the total number of the expansion coefficients of the unknown field (uiu_{i} or viv_{i}). This number is also equal to the total number of linear algebraic equations in the system. Table 1 lists the level of refinement and the corresponding DOF.

Table 1: The degree of freedom of FEM for different resolutions
Refinement DOF
252^{5} 202818
262^{6} 1597570
272^{7} 12681474

Figure 2 shows the waveform observed by an observer located at r=46.63Secr=46.63{\rm Sec}. Dashed lines represent the numerical results and the solid lines are obtained from Eq. (29). The dashed vertical line indicates the epoch that the wavefront just arrives at the observer. Dashed lines with different colors show the results with different resolutions. With high spatial resolution 272^{7} (blue dashed line), the numerical results agree with the analytic one very well. This is expected as the accuracy of the FEM is strongly dependent on the resolution of elements.

After investigating the time-domain waveform for a particular local observer, we consider the integral property of the wave equation. Inserting Eq. (29) into Eq. (12), we obtain the total energy of waves in the domain

E(t)=πRi{ωRisin[2ω(tRi)]cos[2ω(tRi)]+1}+2πω2Ri2(tRi),t[Ri,Ro].\displaystyle\begin{split}E(t)=&\pi R_{i}\left\{\omega R_{i}\sin[2\omega(t-R_{i})]-\cos[2\omega(t-R_{i})]+1\right\}\\ +&2\pi\omega^{2}R_{i}^{2}(t-R_{i}),\quad t\in[R_{i},R_{o}]\,.\end{split} (31)

Then we measure the total energy in our simulation numerically.

Figure 3 compares the numerical results (stars) with the analytical ones (solid line) from Eq. (31). When the wavefront has not reached the outer boundary, namely, Ri<t<RoR_{i}<t<R_{o}, the numerical results agree with the analytic ones very well. After t>Rot>R_{o} (vertical dashed line), the wavefront starts getting out of the simulation domain. The total energy inside the bulk of our simulation domain becomes steady (flat) as, in this case, the energy getting into the simulation domain equals the energy getting out of the domain.

Refer to caption
Figure 3: Upper panel: Comparisons of the total energy between numerical simulations (stars) and the exact analytic solution (solid line). Red and blue colors are for spatial resolution 262^{6} and 272^{7}, respectively. The dashed vertical line indicates the epoch that the wavefront arrives at the outer boundary of our simulation domain. After t>Rot>R_{o}, waves start propagating out of the simulation domain. Lower Panel: relative errors between the numerical solutions and the exact analytic solution. In FEM, spatial resolution is the main driver of errors. By increasing the spatial resolution, the errors can be reduced significantly.

5 Gravitational Waves in a potential well

In this section, we test the accuracy and robustness of our code in simulating GWs traveling in a potential well. We first introduce the basic equations and then discuss the numerical setups of solving these equations.

5.1 Basic equations

We assume a metric gμνg_{\mu\nu} given by

ds2=(1+2ψ)dt2+(12ψ)dx2,ds^{2}=-(1+2\psi)dt^{2}+(1-2\psi)dx^{2}\,, (32)

and the gravitational waves as a perturbation to the background metric

gμν=gμν0+hμν,g_{\mu\nu}=g_{\mu\nu}^{0}+h_{\mu\nu}\quad, (33)

where gμν0g_{\mu\nu}^{0} is the background metric. Following Ref. Peters (1974), we neglect higher order non-linear terms and arrive at the propagation equation for gravitational waves hijh_{ij}

2hij(14ψ)2t2hij=0.\nabla^{2}h_{ij}-(1-4\psi)\frac{\partial^{2}}{\partial t^{2}}h_{ij}=0\,. (34)

Using the eikonal approximation by Ref. (Baraldo et al. (1999)) (also see Ref. Weinberg (1972)), the gravitational wave tensor can be represented as

hij=heij,h_{ij}=he_{ij}\,, (35)

where eije_{ij} is the polarization tensor. In this work, we assume that the polarization tensor does not change during the propagation of GWs. Then we obtain the so-called scalar wave equation

c22h2ht2=0,c^{2}\nabla^{2}h-\frac{\partial^{2}h}{\partial t^{2}}=0\,, (36)

where c2=1/(14ψ)c^{2}=1/(1-4\psi) is the speed of wave.

Unlike the conventional wave equation, one notable feature of this equation is that the wave speed is not a constant but is dependent on the potential well. The varying wave speed plays an important role in the propagation of GWs in a potential well. Moreover, given the fact that ψ<0\psi<0, we have c<1c<1. The presence of potential thus delays the propagation of GWs. This phenomenon is known as the Shapiro time-delay Shapiro (1964).

5.2 Plane wave approximation at the simulation box

Figure 4 shows the domain and setups of our simulations, where RiR_{i} is the distance from the source to the surface of the simulation domain. We assume that the source of GWs is far away from the simulation domain. The incident wave train travels along the zz-axis and enters the surfaces of the simulation domain at z=0z=0. Due to the long-range nature of gravity produced by the scatterer, the incident waves suffer the Shapiro time delay on their way from the source to the scatterer, which adds up a shift of phase to the wave function at the boundary of the simulation domain. Since the source is distant from the scatterer and the length of the simulation domain is small relative to RiR_{i}, the wave function at the simulation domain can be well approximated by plane waves with a constant phase shift Δt=2Mln(2Ri/d)\Delta t=-2M\ln(2R_{i}/d), where dd is the distance of the scatterer to the boundary of the simulation domain. Moreover, we choose the size of the simulation domain large enough so that the scatterer does not distort the wave front initially at the boundary of the simulation domain. Such distortion effect can be accurately and robustly determined using the null geodesic equations, which are provided in appendix B. After these considerations, we set the boundary condition at z=0z=0 as

{h(t)|z=0=Q(tRi/c)4πRih˙(t)|z=0=Q˙(tRi/c)4πRi,\displaystyle\left\{\begin{aligned} h(t)|_{z=0}&=\frac{Q(t-R_{i}/c)}{4\pi R_{i}}\,\\ \dot{h}(t)|_{z=0}&=\frac{\dot{Q}(t-R_{i}/c)}{4\pi R_{i}}\,\end{aligned}\right.\,, (37)

where Q(t)Q(t) is the waveform of the source and the dot denotes the derivative of time.

In this subsection, we choose QQ as a sine function. For convenience, the initial phase of the source is so chosen that it cancels the change of phase due to the Shapiro time-delay when waves arrive at the simulation domain. As such, the boundary condition at z=0z=0 can be represented as

{h(t)|z=0=sin(ωt)h˙(t)|z=0=ωcos(ωt),\displaystyle\left\{\begin{aligned} h(t)|_{z=0}&=\sin(\omega t)\,\\ \dot{h}(t)|_{z=0}&=\omega\cos(\omega t)\,\end{aligned}\right.\,, (38)

where ω=2πf\omega=2\pi f and ff is the frequency of the wave. The amplitude of the incident wave is normalized as unity at z=0z=0. Moreover, we choose the zero point of the temporal axis at the epoch of the wavefront arriving at the surface of the simulation box. For other surfaces of the simulation domain, we adopt the absorbing boundary conditions as explained in Eq. (27). The absorbing boundary conditions aim to guarantee that GWs hh do not reflect into the simulation box when they arrive at the boundaries. In addition to the boundary conditions, we set the initial condition inside the simulation domain as

h=0|t=0,h=0|_{t=0}\,,

which means that there are no waves in the simulation domain initially.

Refer to caption
Figure 4: The domain of our simulations. RiR_{i} is the distance from the origin of the simulation domain to the source of the GWs.

For ψ\psi, we choose it as the potential generated by a solid spherical ball with uniform density

ψ={Mxx0xx0>RsM3Rs2xx022Rs3xx0Rs,\psi=\left\{\begin{aligned} &-\frac{M}{\left\|x-x_{0}\right\|}\quad&\left\|x-x_{0}\right\|>R_{s}\\ &-M\frac{3R_{s}^{2}-\left\|x-x_{0}\right\|^{2}}{2R_{s}^{3}}\quad&\left\|x-x_{0}\right\|\leq R_{s}\end{aligned}\right.\,, (39)

where Rs=2MR_{s}=2M is the Schwarzschild radius and x0x_{0} is the center of the potential well. Note that, ψ\psi is continuous up to the first order of derivative, namely, ψC1(Ω)H1(Ω)\psi\in C^{1}(\Omega)\subset H^{1}(\Omega). We choose the mass of the scatterer as

M=105M.M=10^{5}M_{\odot}\,. (40)
Refer to caption
Figure 5: The waveform of a finite wave train in a potential well. From left to right, the panels show waveforms at different times. The waves enter the simulation domain at z=0z=0 and then travels normally along the zz-axis. The waveform is a sine function but only lasts for one period. The snapshots are taken along the zyz-y plane at x=0x=0. The color bar to the right shows the amplitude of waves hh. The black dots indicate the position of the scatterer. As the wave speed is dependent on the potential well c2=1/(14ψ)c^{2}=1/(1-4\psi), waves travel faster far away from the center than those close to the center. The leading wavefront is bent toward the center (left panel). Unlike the thin-lens model as in the literature, after the scatterer, the deflection of the wavefront is so large that the wavefront can get warped, which leads to a complex 3D shape of the wave zone even when the waves get out of the potential well of the scatterer (right panel).

In this work, we first simulate the behavior of the leading wavefront for a finite wave train in a potential well, we choose the input wave train as a sine function but only lasting for one period

h(t)|z=0={sin(ωt),tλ0,t>λ,\displaystyle h(t)|_{z=0}=\left\{\begin{aligned} &\sin(\omega t)\,,&t\leq\lambda\\ &0\,,&t>\lambda\end{aligned}\right.\,, (41)

where λ\lambda is the period. We choose the simulation domain as a cube. We choose the refinement of the simulation domain along each dimension in 3D space as 292^{9}. The size of the simulation box is taken as 20[sec]20\,[\rm sec] along one side. We place the scatterer at the center of the box. As shown in appendix B, in this case the scatterer is distant from the boundary surface of the incident waves. As such, the distortion effect of the scatterer at boundaries is negligible.

Figure 5 shows the waveform of the finite wave train in a potential well. We choose λ=4[sec]\lambda=4[\rm sec] for illustrative purposes. From left to right, the panels show waveforms at different times. The snapshots are taken along the zyz-y plane at x=0x=0. The color bar to the right shows the amplitude of waves hh. The black dot indicates the position of the scatterer. As the wave speed is dependent on the potential well c2=1/(14ψ)c^{2}=1/(1-4\psi), waves travel faster far away from the center than those closer to the center. The leading wavefront is bent toward the center (left panel). After the scatterer, the wavefront is warped, which leads to a complex 3D shape of the wave zones even when the waves get out of the potential well of the scatterer (right panel).

5.3 Numerical results against analytical predictions on the wavefront

Unlike the rest part of waves, the leading wavefront represents the transport of energy from one place to another. Therefore, it is strictly subject to the constraint of causality. Moreover, the hypersurface of the wavefront (denoted as SS) is a null hypersurface. Its normal vector ka=aSk^{a}=\nabla^{a}S is a null vector kaka=0k^{a}k_{a}=0. As shown in Wald (2010) (see, Eq.(4.2.37)), the integral curves of kak^{a} are null geodesics, which satisfy kaakb=0k^{a}\nabla_{a}k^{b}=0 if kak^{a} is presented in terms of an affine parameter. As such, the behavior of the leading wavefront can be traced by null geodesics, which provide an independent way to test our numerical results. As shown in the appendix, the equations for null geodesics can be presented as (see appendix B for details)

d2μdφ2+μ=12μ[μ2ζ(M,μ)],\displaystyle\frac{d^{2}\mu}{d\varphi^{2}}+\mu=-\frac{1}{2}\frac{\partial}{\partial\mu}\left[\mu^{2}\zeta(M,\mu)\right]\,, (42)

where μ=1/r\mu=1/r and rr is the distance to the center of the mass. φ\varphi is the azimuthal angle in the spherical coordinates. The definition of ζ(M,μ)\zeta(M,\mu) is provided in the appendix B. As the incident wave is a plane wave at z=0z=0, the spacial components of its normal vector kak^{a} are parallel to the zz-axis. Therefore, we set the initial conditions for Eq. (42) at z=0z=0 as

μ(φ)1bsinφ,\displaystyle\mu(\varphi)\rightarrow\frac{1}{b}\sin\varphi\,, (43)

where bb is the impact factor. Eq. (43) simply describes a straight line in spherical coordinates along the zz-axis, which is also the solution of Eq. (42) in the absence of the scatterer ζ(M,μ)=0\zeta(M,\mu)=0.

Refer to caption
Figure 6: The integral curves of the normal vectors of the wavefront(the bundle of thin solid black lines). These curves follow the equation of null geodesics. The wavefront predicted by the null-geodesics (red solid line) agrees very well with simulations (colour shaded regions). Moreover, in the thin-lens regimes that are far away from the scatterer, our numerical results also agree with the predictions of the Shapiro-time delay (blue solid line) very well.

Figure 6 shows the trajectories of the null geodesics of the normal vectors kak^{a} of the wavefront obtained using Eq. (42) (the bundle of thin solid black lines). The wavefront predicted by null geodesics at t=17.40[sec]t=17.40[{\rm sec}] (solid red lines) agrees very well with those of simulations (shaded regions). Using null geodesics, it is also clear that after the scatterer, the wavefront get warped and the signals of the leading wavefront on the zz-axis does not come from the signals traveling on the zz-axis itself, but from the wavefront circling around it.

In addition to the tests using null geodesics, we also test our numerical results using the analytic expressions of the Shapiro-time delay. To to this, we introduce a new notion

tShapiro=0tcdt0t(1+2ψ)=tΔtShapiro,\displaystyle t_{\rm Shapiro}=\int_{0}^{t}c\,\mathrm{d}t\approx\int_{0}^{t}(1+2\psi)=t-\Delta t_{\rm Shapiro}\,, (44)

where cc is the speed of wave c=1/(14ψ)c=\sqrt{1/(1-4\psi)}. For a point mass, the Shapiro time delay ΔtShapiro\Delta t_{\rm Shapiro} is given by

ΔtShapiro=2ψ𝑑z=2Mzizfdzz2+b2,\displaystyle\Delta t_{\rm Shapiro}=-2\int\psi dz=2M\int_{z_{i}}^{z_{f}}\frac{dz}{\sqrt{z^{2}+b^{2}}}\,, (45)

where bb is the impact factor and the integration is taken along the zz direction under the thin-lens assumption. zi{z_{i}} and zf{z_{f}} are the initial and final positions of the integration limits. In this work, we take zi=10[sec]z_{i}=-10[{\rm sec}] and zf=7.40[sec]{z_{f}}=7.40[{\rm sec}] which is the zz component of the position of the wavefront relative to the scatterer without delay at time t=17.40[sec]t=17.40[{\rm sec}]. Fig. 6 shows that in regimes that are far away from the scatterer (where the thin-lens assumption is valid), our numerical results agree with the predictions of the Shapiro-time delay (blue solid line) very well.

Refer to caption
Figure 7: Numerical convergence tests for waveforms for simulations with different resolutions. The dashed and solid lines are for simulations with 282^{8} and 292^{9} refinements along each dimension in 3D space, respectively. The solid black vertical line indicates the position of the scatterer. The overall waveforms for simulations with 282^{8} and 292^{9} refinements agree very well. However, a high-resolution run can suppress the wavefront shocks, which are numerical artifacts.

Moreover, we demonstrate the robustness of our numerical results. To do this, we compare the results obtained from the runs with refinements of 282^{8} and 292^{9} for wave functions at different times along the zz-axis. From Fig. 7, the waveforms for simulations with 282^{8} (dashed lines) agree with 292^{9} refinements (solid lines) very well, which means that even with 282^{8} refinement, it already yields good accuracy. However, a high-resolution simulation, such as 292^{9} refinement, can suppress the wavefront shocks, which are numerical artifacts appearing in the wave functions after the scatterer.

5.4 Numerical results against analytical predictions on waveform

After testing the leading wavefront, we further test the waveform of our simulations against analytical predictions. In this new test, we take the incident wave function as a continuous sine function. Moreover, since the incident waves are of azimuthal symmetry, we choose the simulation domain as a cylinder instead of a cube to minimize the effects of boundaries. The incident waves are assumed to travel along the axis of the cylinder(zz axis). The length of the cylinder is taken as 70[sec]70[\rm sec] and the radius is taken as 25[sec]25[\rm sec]. To make it easier to be compared with analytical predictions, we simulate the waves with the evolution time long enough that the waves in the simulation domain can reach a steady state. In this case, the waveform can be analyzed in the Fourier space.

h(x,t)=h(x)eiωt,\displaystyle h(\vec{x},t)=h(\vec{x})e^{-i\omega t}\,, (46)

where ω=2πf\omega=2\pi f is the angular frequency. Equation (36) then becomes

(2+ω2)h=4ω2ψh.\displaystyle(\nabla^{2}+\omega^{2})h=4\omega^{2}\psi h\,. (47)

Compared with Eq. (36), the potential ψ\psi appears as a source term in Eq. (47). As such, the wave speed in Eq. (47) is unity and the wave number k~\tilde{k} equals to the angular frequency k~=ω\tilde{k}=\omega, which is different from that in Eq. (36).

Outside the Schwarzschild radius of the scatterer, the potential ψ\psi of the scatterer can be described by M/r-M/r. In this case, Eq. (47) can be solved analytically(e.g.  Peters (1974)). However, unlike Eq. (36) which is hyperbolic, Eq. (47) is elliptic, which are specific to the “boundary-value" problems Grisvard (1985). The solutions of Eq. (47) are strongly dependent on the boundary conditions imposed. Since the wave functions are of azimuthal symmetry, we impose the following ansatz as boundaries

h~=eik~zf(ξ),\displaystyle\tilde{h}=e^{i\tilde{k}z}f(\xi)\,, (48)

where ξ=rz\xi=r-z and r=xx0=x2+y2+z2r=\left\|x-x_{0}\right\|=\sqrt{x^{2}+y^{2}+z^{2}}(the scatterer is placed at the origin). Since f(ξ)f(\xi) does not depend on the direction of k~\tilde{k}, h~\tilde{h} only has Fourier modes with wave vectors k~\tilde{k} along the zz-axis. As such, the above ansatz implicitly assumes that the wave vector k~\tilde{k} of the outgoing waves at infinity is still along the zz-axis. Under such assumption, the solution of Eq. (47) is given by

h~=Ceik~zF11(iωM,1,iωξ),\displaystyle\tilde{h}=Ce^{i\tilde{k}z}{{}_{1}F_{1}}(i\omega M,1,i\omega\xi)\,, (49)

where F11{{}_{1}F_{1}} is the confluent hypergeometric function and CC is a complex constant (detailed discussions are provided in Appendix C). When M=0M=0

F11(0,1,iωξ)=1.\displaystyle{{}_{1}F_{1}}(0,1,i\omega\xi)=1\,.

Wave function h~\tilde{h} simply goes back to the plane wave. Moreover, it is worth noting that h~\tilde{h} can not describe waveform on the zz-axis because F11{{}_{1}F_{1}} is not well defined when ξ=0\xi=0.

If rRsr\gg R_{s} and ξ1\xi\gg 1, the confluent hypergeometric F11(iωM,1,iωξ){{}_{1}F_{1}}(i\omega M,1,i\omega\xi) can be expanded around iωξ=i\omega\xi=\infty as

F11(iωM,1,iωξ)ξiωM[1+ω2M2ξ].\displaystyle{{}_{1}F_{1}}(i\omega M,1,i\omega\xi)\approx\xi^{-i\omega M}\left[1+\frac{\omega^{2}M^{2}}{\xi}\right]\,. (50)

In this case, the wave function can be written as

h~(x,y,z)Ceik~z(x2+y2+z2z)iωM[1+ω2M2x2+y2+z2z].\displaystyle\tilde{h}(x,y,z)\approx Ce^{i\tilde{k}z}(\sqrt{x^{2}+y^{2}+z^{2}}-z)^{-i\omega M}\left[1+\frac{\omega^{2}M^{2}}{\sqrt{x^{2}+y^{2}+z^{2}}-z}\right]\,. (51)

Note that h~(x,y,z)\tilde{h}(x,y,z) are complex values. The spatial waveform can then be obtained by taking the real part of h~(x,y,z)\tilde{h}(x,y,z). The complex constant CC is independent of (x,y,z)(x,y,z), which can be determined by matching the amplitude and phase of h~(x,y,z)\tilde{h}(x,y,z) with the incident waves. An advantage of the above equation is that it can be evaluated efficiently in the numerical processes. We therefore use it to test our simulation results.

Refer to caption
Figure 8: The waveform of a continuous wave train in a potential well. The snapshot is taken along the zyz-y plane with x=0x=0 at t=136.50[Sec]t=136.50[\rm Sec]. Since the incident waves are of azimuthal symmetry, in this case the simulation domain is taken as a cylinder with the incident waves traveling along the direction of the axis of the cylinder(zz-axis). The wavefrom of the incident waves is a continuous sine function with a frequency of f=0.25[Hz]f=0.25\,[{\rm Hz}]. The time evolution is taken long enough so that the waveform is in a steady state. The black dot indicates the position of the scatterer. The signals after the scatterer show a clear diffraction pattern with strong signals along the zz-axis.

Figure 8 shows a snapshot of our simulation along the zyz-y plane at x=0x=0. The incident wave travels along the zz-axis with a waveform as a continuous sine function. The frequency of the incident wave is f=0.25[Hz]f=0.25\,[{\rm Hz}]. The snapshot is taken at t=136.50[Sec]t=136.50[\rm Sec], for which the time evolution is long enough so that the waveform is in a steady state. We have checked that further increasing the evolution time does not change the waveform appreciably. The black dot indicates the position of the scatterer. The signals after the scatterer show a clear diffraction pattern with strong signals along the zz-axis.

Refer to caption
Figure 9: The 1D waveform of simulations against analytical predictions. The waveforms are taken along different lines that are parallel to the zz-axis. The distances of these lines to the zz-axis are taken as y=10.0563[Sec]y=10.0563[\rm Sec], y=14.0230[Sec]y=14.0230[\rm Sec] and y=18.0946[Sec]y=18.0946[\rm Sec], respectively. The black lines show the waveform from simulations and the red lines show the analytical predictions. The simulations agree with the analytical predictions very well.

Figure 9 compares the wavefroms of our simulations (black lines) with the analytical predictions (red lines). The wavefroms are taken along 1D lines that are parallel to the zz-axis. Since the wave function and simulation domain are of azimuthal symmetry, we only extract these lines from the zyz-y plane. The distances of these lines to the zz-axis are chosen as y=10.0563[Sec]y=10.0563[\rm Sec], y=14.0230[Sec]y=14.0230[\rm Sec] and y=18.0946[Sec]y=18.0946[\rm Sec], respectively. The waveform of our simulations agree with the analytical predictions very well.

6 Conclusions

The detection of gravitational waves in the low-frequency regime provides a powerful tool to probe the properties of supermassive back holes (SMBHs) throughout cosmic history, which has profound implications in gravitational physics and galaxy formation physics. However, due to the wave nature of GWs, wave effects such as diffraction or interference may hamper the future GW experiments to accurately infer the source information from the observed signals. Accurately pinning down such effects, therefore, is urgently called for.

The wave effects of GWs have already been discussed in the literature. However, the previous analyses focus on the frequency-domain, coupled to some implicit assumptions, such as that GWs are in a steady-state (e.g. Ref. Peters (1974)) or stationary state (e.g. Ref. Suyama et al. (2005)). These analyses also assume that the wave zone of GWs is infinite, which neglects some important wave effects such as the causality and wavefront effects. For a time-finite GW signal, these two effects may have significant impacts on the waveforms. Contrary to common intuition that the time-domain and frequency-domain are equivalent, the wavefront effects in a potential well should be studied in the time-domain. This is because the propagation of wavefronts is a “Cauchy" problem for hyperbolic equations, which is defined in part of the spacetime and is fundamentally local (more discussions can be found in section 10.1 of Wald (2010)). If taking the Fourier transforms, the wave equations in the frequency-domain become elliptic (e.g. Helmholtz equations), which are specific to the “boundary-value" problems Grisvard (1985). The wave functions, in this case, are no longer determined by initial conditions but by boundary conditions. In this case, the boundaries in the future at infinity have to be specified, which can only be consistently set for simple cases, such as waves traveling in free space. However, if for a complicated problem, such as waves traveling in a potential well, the wave-zones may have complex shapes in 3D space and it is difficult to set consistent boundary conditions. In these cases, GWs should be studied in the time-domain directly.

Motivated by these facts, in this paper, based on the publicly available code deal.ii as well as stand-alone linear algebra libraries, such as PETSc, we have developed a code called GWsim to simulate the propagation of GWs in a potential well. We have tested our code using a monochromatic spherical wave from a point source. We find that not only for the waveforms at an individual observer but also for the conservation of energy, our numerical results agree with the analytical predictions very well (see Fig. 3).

After testing our code, we have studied the propagation of wavefront in a potential well. We focus particularly on a time-domain specific problem, namely, the leading wavefront in a potential well. Since the integral curves of the normal vectors of the leading wavefront are null-geodesics, the leading wavefront can be traced by the null geodesics, which provide an independent way to test our numerical results. We find that the wavefront predicted by null geodesics agree with our numerical results very well (see Fig. 6). In addition to the null geodesics, we also compared our numerical results with the analytical expressions of the Shapiro-time delay

ΔtShapiro=2ψ𝑑z=2Mzizfdzz2+b2.\displaystyle\Delta t_{\rm Shapiro}=-2\int\psi dz=2M\int_{z_{i}}^{z_{f}}\frac{dz}{\sqrt{z^{2}+b^{2}}}\,. (52)

We find that in the thin-lens regimes, namely, far away from the scatterer, our numerical results agree with the predictions of Eq. (52) very well. However, near the optical axis, we find that Eq. (52) can not give accurate predictions of the positions of the leading wavefront. The actual wavefront travels faster than the predictions of Eq. (52). These results suggest that the analyses presented in Suyama (2020) and Ezquiaga et al. (2020) is only valid in the thin-lens regimes since the authors there use Eq. (52) to predict the speed of GWs. Moreover, it is also worth noting that unlike the conventional images in geometric optics, GWs can not be sheltered by the scatterer due to wave effects. The signals of GWs can circle around the scatterer and travel along the optic axis until arrive at a distant observer(see e.g. Fig. 6), which is an important observational consequence of GWs in such a system.

In addition to the above conclusions, due to the linearity of the wave equation and the geometric unit, our numerical results can serve as templates. The mass of the scatterer used in this work can be rescaled to other values that are of astrophysical interest. For instance, if we want to obtain the results for a scatterer with a mass of M=107MM=10^{7}M_{\odot}, we only need to re-scale the temporal and spatial axes of our numerical results by 100100 times.

Finally, this work is the first attempt to simulate the wave effects of GWs in the time domain. Therefore, following the previous work in the literature Deguchi & Watson (1986a, b); Schneider et al. (1992); Ruffa (1999); De Paolis et al. (2002); Takahashi & Nakamura (2003); Nakamura & Deguchi (1999); Suyama et al. (2005); Christian et al. (2018); Zakharov & Baryshev (2002); Liao et al. (2019); Macquart (2004); Nakamura (1998); Dai et al. (2018); Cao et al. (2014); Yoo et al. (2013); Nambu et al. (2019), we treat the potential well of the scatterer as Newtonian gravity, which is reasonable in this work, since we only focus on the leading wavefront, for which most signals come from regimes that are far away from the scatterer. However, to fully explore the effects of general relativity on GWs requires new development of the current code, which is beyond the scope of this paper. In a series of follow-up papers, we will extend our work to the spacetime of a realistic black hole, such as a Kerr black hole Baraldo et al. (1999). The method presented in this paper will be a powerful tool for exploring GWs in a non-static potential well, which is of significant astrophysical interests. Moreover, we will also explore the effectiveness of using the propagation of GWs as a tool to test the theory of gravity Chesler & Loeb (2017); Belgacem et al. (2019); Koyama (2020).

Acknowledgements

J.H.H. acknowledges support of Nanjing University. This work is supported by the National Natural Science Foundation of China (12075116). The numerical calculations in this paper have been done on the computing facilities in the High Performance Computing Center (HPCC) of Nanjing University.

Data Availability

The data and code underlying this article will be shared on reasonable request to the corresponding author.

References

  • Abbott et al. (2016) Abbott B. P., et al., 2016, Phys. Rev. Lett., 116, 061102
  • Abhyankar et al. (2018) Abhyankar S., Brown J., Constantinescu E. M., Ghosh D., Smith B. F., Zhang H., 2018, arXiv preprint arXiv:1806.01437
  • Alzetta et al. (2018) Alzetta G., et al., 2018, Journal of Numerical Mathematics, 26, 173
  • Arndt et al. (2019) Arndt D., et al., 2019, Journal of Numerical Mathematics
  • Balay et al. (1997) Balay S., Gropp W. D., McInnes L. C., Smith B. F., 1997, in Arge E., Bruaset A. M., Langtangen H. P., eds, Modern Software Tools in Scientific Computing. Birkhäuser Press, pp 163–202
  • Balay et al. (2019a) Balay S., et al., 2019a, PETSc Web page, https://www.mcs.anl.gov/petsc, https://www.mcs.anl.gov/petsc
  • Balay et al. (2019b) Balay S., et al., 2019b, Technical Report ANL-95/11 - Revision 3.12, PETSc Users Manual, https://www.mcs.anl.gov/petsc. Argonne National Laboratory, https://www.mcs.anl.gov/petsc
  • Bangerth et al. (2007) Bangerth W., Hartmann R., Kanschat G., 2007, ACM Trans. Math. Softw., 33, 24/1
  • Baraldo et al. (1999) Baraldo C., Hosoya A., Nakamura T. T., 1999, Phys. Rev., D59, 083001
  • Bartos et al. (2017) Bartos I., Kocsis B., Haiman Z., Márka S., 2017, Astrophys. J., 835, 165
  • Belgacem et al. (2019) Belgacem E., et al., 2019, JCAP, 1907, 024
  • Bellovary et al. (2016) Bellovary J. M., Mac Low M.-M., McKernan B., Ford K. E. S., 2016, Astrophys. J., 819, L17
  • Bontz & Haugan (1981) Bontz R. J., Haugan M. P., 1981, apss, 78, 199
  • Cao et al. (2014) Cao Z., Li L.-F., Wang Y., 2014, Phys. Rev. D, 90, 062003
  • Chesler & Loeb (2017) Chesler P. M., Loeb A., 2017, Phys. Rev. Lett., 119, 031102
  • Christian et al. (2018) Christian P., Vitale S., Loeb A., 2018, Phys. Rev., D98, 103022
  • Dai et al. (2018) Dai L., Li S.-S., Zackay B., Mao S., Lu Y., 2018, Phys. Rev. D, 98, 104029
  • De Paolis et al. (2002) De Paolis F., Ingrosso G., Nucita A. A., Qadir A., 2002, Astron. Astrophys., 394, 749
  • Deguchi & Watson (1986a) Deguchi S., Watson W. D., 1986a, Phys. Rev. D, 34, 1708
  • Deguchi & Watson (1986b) Deguchi S., Watson W. D., 1986b, Astrophys. J., 307, 30
  • Dwyer et al. (2015) Dwyer S., Sigg D., Ballmer S. W., Barsotti L., Mavalvala N., Evans M., 2015, Phys. Rev. D, 91, 082001
  • Ezquiaga et al. (2020) Ezquiaga J. M., Hu W., Lagos M., 2020, Phys. Rev. D, 102, 023531
  • Ferrarese & Merritt (2000) Ferrarese L., Merritt D., 2000, Astrophys. J., 539, L9
  • Grisvard (1985) Grisvard P., 1985, Elliptic problems in nonsmooth domains. Pitman Advanced Pub. Program Boston
  • Grossmann et al. (2007) Grossmann C., Roos H.-G., Stynes M., 2007, Numerical treatment of partial differential equations. Springer
  • Haehnelt & Kauffmann (2000) Haehnelt M. G., Kauffmann G., 2000, Mon. Not. Roy. Astron. Soc., 318, L35
  • Hobbs et al. (2010) Hobbs G., et al., 2010, Classical and Quantum Gravity, 27, 084013
  • Koyama (2020) Koyama K., 2020, Phys. Rev. D, 102, 021502
  • Lax & Milgram (1954) Lax P. D., Milgram A. N., 1954, in Annals of Mathematics Studies, no. 33, Contributions to the theory of partial differential equations. Princeton University Press, Princeton, N. J., pp 167–190
  • Liao et al. (2019) Liao K., Biesiada M., Fan X.-L., 2019, Astrophys. J., 875, 139
  • Macquart (2004) Macquart J.-P., 2004, Astron. Astrophys., 422, 761
  • Nakamura (1998) Nakamura T. T., 1998, Phys. Rev. Lett., 80, 1138
  • Nakamura & Deguchi (1999) Nakamura T. T., Deguchi S., 1999, Progress of Theoretical Physics Supplement, 133, 137
  • Nambu et al. (2019) Nambu Y., Noda S., Sakai Y., 2019, Phys. Rev. D, 100, 064037
  • Ohanian (1974) Ohanian H. C., 1974, Int. J. Theor. Phys., 9, 425
  • Peters (1974) Peters P. C., 1974, Phys. Rev. D, 9, 2207
  • Punturo et al. (2010) Punturo M., et al., 2010, Classical and Quantum Gravity, 27, 194002
  • Ruffa (1999) Ruffa A. A., 1999, The Astrophysical Journal, 517, L31
  • Sato et al. (2009) Sato S., et al., 2009, Journal of Physics: Conference Series, 154, 012040
  • Schneider et al. (1992) Schneider P., Ehlers J., Falco E. E., 1992, Gravitational Lenses. Springer, doi:10.1007/978-3-662-03758-4
  • Shapiro (1964) Shapiro I. I., 1964, Phys. Rev. Lett., 13, 789
  • Sorge (2015) Sorge F., 2015, Class. Quant. Grav., 32, 035007
  • Stone et al. (2017) Stone N. C., Metzger B. D., Haiman Z., 2017, Mon. Not. Roy. Astron. Soc., 464, 946
  • Suyama (2020) Suyama T., 2020, Astrophys. J., 896, 46
  • Suyama et al. (2005) Suyama T., Takahashi R., Michikoshi S., 2005, Phys. Rev., D72, 043001
  • Takahashi & Nakamura (2003) Takahashi R., Nakamura T., 2003, Astrophys. J., 595, 1039
  • Takahashi et al. (2005) Takahashi R., Suyama T., Michikoshi S., 2005, Astron. Astrophys., 438, L5
  • Wald (2010) Wald R. M., 2010, General relativity. University of Chicago press
  • Weinberg (1972) Weinberg S., 1972, Gravitation and cosmology: principles and applications of the general theory of relativity. John Wilry and Sons, Inc.
  • Yoo et al. (2013) Yoo C.-M., Harada T., Tsukamoto N., 2013, Phys. Rev. D, 87, 084045
  • Young & Chen (1996) Young D. M., Chen J. Y., 1996, doi:10.2172/409863
  • Zakharov & Baryshev (2002) Zakharov A. F., Baryshev Y. V., 2002, Classical and Quantum Gravity, 19, 1361
  • Zienkiewicz & Taylor (1989) Zienkiewicz O., Taylor R., 1989, The finite element method. Vol.1, Basic formulation and linear problems. London:McGraw-Hill
  • eLISA Consortium et al. (2013) eLISA Consortium et al., 2013, arXiv e-prints, p. arXiv:1305.5720

Appendix A Spherical Wave in free-space

We adopt the following convention for the Fourier transform

u(r,θ,ϕ,t)\displaystyle u(r,\theta,\phi,t) =u(r,θ,ϕ,ω)eiωtdt,\displaystyle=\int u(r,\theta,\phi,\omega)e^{-i\omega t}\,\mathrm{d}t\,,
u(r,θ,ϕ,ω)\displaystyle u(r,\theta,\phi,\omega) =12πu(r,θ,ϕ,t)eiωtdt.\displaystyle=\frac{1}{2\pi}\int u(r,\theta,\phi,t)e^{i\omega t}\,\mathrm{d}t\,.

In the frequency-domain, Eq. (1) in free-space becomes

[2+(ω/c)2]u=0.\left[\nabla^{2}+(\omega/c)^{2}\right]u=0\,. (53)

For a diverging spherical wave radiating from the origin, the general solution of the above equation has the form

u(r,θ,ϕ,ω)=lmalm(ω)hl(1)(ωcr)Ylm(θ,ϕ),u(r,\theta,\phi,\omega)=\sum_{lm}a_{lm}(\omega)h_{l}^{(1)}\left(\frac{\omega}{c}r\right)Y_{lm}(\theta,\phi)\,, (54)

where hl(1)(x)h_{l}^{(1)}(x) is the spherical Hankel functions of the first kind, which is the solution of the spherical Bessel equation satisfying the radiation boundary condition at infinity. Ylm(θ,ϕ)Y_{lm}(\theta,\phi) is the spherical harmonics. alm(ω)a_{lm}(\omega) are unknown coefficients, which can be determined by the boundary conditions.

We assume that the boundary condition for a spherical wave is given at the surface of a sphere with a radius of RR.

u(r,θ,ϕ,ω)|Ω=fR(θ,ϕ,t),u(r,\theta,\phi,\omega)|_{\partial\Omega}=f_{R}(\theta,\phi,t)\,,

where fR(θ,ϕ,t)f_{R}(\theta,\phi,t) is a known-function. In the frequency-domain, we have

fR(θ,ϕ,ω)=12πfR(θ,ϕ,t)eiωtdt.f_{R}(\theta,\phi,\omega)=\frac{1}{2\pi}\int f_{R}(\theta,\phi,t)e^{i\omega t}\,\mathrm{d}t\,. (55)

We expand fR(θ,ϕ,ω)f_{R}(\theta,\phi,\omega) in terms of spherical harmonics

fR(θ,ϕ,ω)=lmblm(w)Ylm(θ,ϕ).f_{R}(\theta,\phi,\omega)=\sum_{lm}b_{lm}(w)Y_{lm}(\theta,\phi)\,. (56)

Comparing the above equation with Eq. (54) at the boundary, we obtain

alm(ω)hl(1)(ωcR)=blm(ω).a_{lm}(\omega)h_{l}^{(1)}\left(\frac{\omega}{c}R\right)=b_{lm}(\omega)\,. (57)

Thus, we have

u(r,θ,ϕ,t)=eiωtu(r,θ,ϕ,ω)dt=eiωtlmhl(1)(ωcr)hl(1)(ωcR)blm(w)Ylm(θ,ϕ)dt.\displaystyle\begin{split}u(r,\theta,\phi,t)&=\int e^{-i\omega t}u(r,\theta,\phi,\omega)\,\mathrm{d}t\\ &=\int e^{-i\omega t}\sum_{lm}\frac{h_{l}^{(1)}(\frac{\omega}{c}r)}{h_{l}^{(1)}(\frac{\omega}{c}R)}b_{lm}(w)Y_{lm}(\theta,\phi)\,\mathrm{d}t\,.\end{split} (58)

The above equation demonstrates the Huygens-Fresnel principle, namely, the propagation of waves is completely determined by the surface of the wavefront.

If the wavefront at the boundary is of spherical symmetry and is the same as the wavefront propagating there from the point source at the origin

fR(θ,ϕ,t)=Q(tRc)4πR,f_{R}(\theta,\phi,t)=\frac{Q(t-\frac{R}{c})}{4\pi R}\,, (59)

there will be only monopole l=0l=0 left in the expansions of Eq. (58). Further noting that

h0(1)(x)=ixeix,h_{0}^{(1)}(x)=-\frac{i}{x}e^{ix}\,, (60)

Eq. (58) then reduces to

u(r,θ,ϕ,t)=Rreiω(trRc)fR(θ,ϕ,ω)dt=RrfR(θ,ϕ,trRc)=Q(trc)4πr,\displaystyle\begin{split}u(r,\theta,\phi,t)&=\frac{R}{r}\int e^{-i\omega(t-\frac{r-R}{c})}f_{R}(\theta,\phi,\omega)\,\mathrm{d}t\\ &=\frac{R}{r}f_{R}(\theta,\phi,t-\frac{r-R}{c})\\ &=\frac{Q(t-\frac{r}{c})}{4\pi r}\end{split}\,, (61)

which is the same as the wave propagating directly to the radius rr from the origin.

In general cases, if x=ωcr1x=\frac{\omega}{c}r\gg 1

hl(1)(x)1xeix(i)l+1,h_{l}^{(1)}(x)\sim\frac{1}{x}e^{ix}(-i)^{l+1}\,,

we have

hl(1)(ωcr)hl(1)(ωcR)Rreiωc(rR).\displaystyle\frac{h_{l}^{(1)}(\frac{\omega}{c}r)}{h_{l}^{(1)}(\frac{\omega}{c}R)}\sim\frac{R}{r}e^{i\frac{\omega}{c}(r-R)}\,. (62)

The wave function reduces to

u(r,θ,ϕ,t)=Rreiω(trRc)fR(θ,ϕ,ω)dt=RrfR(θ,ϕ,trRc).\displaystyle\begin{split}u(r,\theta,\phi,t)&=\frac{R}{r}\int e^{-i\omega(t-\frac{r-R}{c})}f_{R}(\theta,\phi,\omega)\,\mathrm{d}t\\ &=\frac{R}{r}f_{R}(\theta,\phi,t-\frac{r-R}{c})\end{split}\,. (63)

This indicates that the anisotropic feature in the wave function at the radius RR can be preserved to a distant observer during the propagation of waves, if the waves travel in a free-space. The amplitude of the wave decreases by a factor of Rr\frac{R}{r}, which is the same as the isotropic wave, which is due to the conservation of energy of waves.

Appendix B Null geodesics in curved spacetime

Refer to caption
Figure 10: The right-hand side of Eq.( 74) as a function of radius rr. For r2Mr\gg 2M, μ[μ2ζ(M,μ)]\frac{\partial}{\partial\mu}\left[\mu^{2}\zeta(M,\mu)\right] decays as 1/r21/r^{2}, which is much faster than that of the potential.
Refer to caption
Figure 11: The trajectories of null geodesics (black solid lines) in the spacetime of the scatterer. The center of the scatterer is at z=10[sec]z=10[\rm sec]. The initial conditions of Eq. (75) for the null geodecis equations are set at z=40[sec]z=-40[\rm sec], 50[sec]50[\rm sec] away from the scatterer. The red curves indicate the wavefront at different killing times tt. As Eq. (74) can not be applied to the zz-axis, we do not show light rays on the zz-axis. The impact of the scatterer on wavefront is only significant when the wavefront is close to the scatterer within a radius of 10[sec]10[\rm sec]. However, outside this region, the distortion effect on the wavefront becomes negligible.

In this subsection, we describe the null geodesics in the spacetime of a point mass. We assume that the line element has a generic form

ds2=g00(M,r)dt2+g11(M,r)dr2+r2(dθ2+sinθ2dφ2),\displaystyle ds^{2}=g_{00}(M,r)dt^{2}+g_{11}(M,r)dr^{2}+r^{2}(d\theta^{2}+\sin\theta^{2}d\varphi^{2})\,, (64)

where g00(M,r)g_{00}(M,r) and g11(M,r)g_{11}(M,r) are functions of mass and radius from the point mass. For a null-geodesic, (τ)a\left(\frac{\partial}{\partial\tau}\right)^{a} denotes the tangent vector and τ\tau is the affine parameter. We assume that the initial position and the tangent vector of the geodesic lie in the “equatorial plane" θ=π/2\theta=\pi/2 (this is always possible by choosing proper coordinates). The null condition gab(τ)a(τ)b=0g_{ab}\left(\frac{\partial}{\partial\tau}\right)^{a}\left(\frac{\partial}{\partial\tau}\right)^{b}=0 gives

g00(M,r)(dtdτ)2+g11(M,r)(drdτ)2+r2(dφdτ)2=0.\displaystyle g_{00}(M,r)\left(\frac{dt}{d\tau}\right)^{2}+g_{11}(M,r)\left(\frac{dr}{d\tau}\right)^{2}+r^{2}\left(\frac{d\varphi}{d\tau}\right)^{2}=0\,. (65)

As (t)a\left(\frac{\partial}{\partial t}\right)^{a} is a static time-like killing vector, the quantity

E=gab(τ)a(t)b=g00(dtdτ)\displaystyle E=-g_{ab}\left(\frac{\partial}{\partial\tau}\right)^{a}\left(\frac{\partial}{\partial t}\right)^{b}=-g_{00}\left(\frac{dt}{d\tau}\right)\, (66)

along the geodesic (τ)a\left(\frac{\partial}{\partial\tau}\right)^{a} is a constant. Likewise, for the rotational Killing vector (φ)a(\frac{\partial}{\partial\varphi})^{a}

L=r2(dφdτ)\displaystyle L=r^{2}\left(\frac{d\varphi}{d\tau}\right)\, (67)

is a constant. Inserting EE and LL back to Eq. (65), we obtain

g001(M,r)E2+g11(M,r)(drdτ)2+L2r2=0.\displaystyle g_{00}^{-1}(M,r)E^{2}+g_{11}(M,r)\left(\frac{dr}{d\tau}\right)^{2}+\frac{L^{2}}{r^{2}}=0\,. (68)

Further by noting that

drdτ=(drdφ)Lr2,\displaystyle\frac{dr}{d\tau}=\left(\frac{dr}{d\varphi}\right)\frac{L}{r^{2}}\,, (69)

and using the condition

g00(M,r)g11(M,r)=1,\displaystyle g_{00}(M,r)g_{11}(M,r)=-1\,, (70)

we arrive at

(drdφ)2E2r4L2+r2g11=0.\displaystyle\left(\frac{dr}{d\varphi}\right)^{2}-\frac{E^{2}r^{4}}{L^{2}}+\frac{r^{2}}{g_{11}}=0\,. (71)

If we denote μ=1/r\mu=1/r, the above equation can be cast into

(dμdφ)2=E2L2μ2g11.\displaystyle\left(\frac{d\mu}{d\varphi}\right)^{2}=\frac{E^{2}}{L^{2}}-\frac{\mu^{2}}{g_{11}}\,. (72)

It is more convenient to write g111(M,r)g_{11}^{-1}(M,r) as

g111(M,r)=1+ζ(M,μ).\displaystyle g_{11}^{-1}(M,r)=1+\zeta(M,\mu)\,. (73)

Taking the derivative of Eq. (72) and noting that EE and LL are constants, we obtain

d2μdφ2+μ=12μ[μ2ζ(M,μ)].\displaystyle\frac{d^{2}\mu}{d\varphi^{2}}+\mu=-\frac{1}{2}\frac{\partial}{\partial\mu}\left[\mu^{2}\zeta(M,\mu)\right]\,. (74)

The above equation gives the trajectories of null geodesics in terms of μ\mu and φ\varphi. However, to solve Eq. (74), we need to set the initial conditions. We assume that the source of the incident light rays are far away from the point mass and are parallel to the zz-axis. In this case, ζ(M,μ)0\zeta(M,\mu)\rightarrow 0 and Eq. (74) has the solution

μ(φ)1bsinφ,\displaystyle\mu(\varphi)\rightarrow\frac{1}{b}\sin\varphi\,, (75)

where bb is the impact factor. The above equation is simply a straight line in the spherical coordinates. We use Eq. (75) as initial conditions at the boundary of the simulation box. Comparing to Eq. (32), we find

ζ(M,μ)=2ψ(M,μ)12ψ(M,μ),\displaystyle\zeta(M,\mu)=\frac{2\psi(M,\mu)}{1-2\psi(M,\mu)}\,, (76)

and

μ[μ2ζ(M,μ)]=8μ2M(Mμ+3/4)(2Mμ+1)2.\displaystyle\frac{\partial}{\partial\mu}\left[\mu^{2}\zeta(M,\mu)\right]=-\frac{8\mu^{2}M(M\mu+3/4)}{(2M\mu+1)^{2}}\,. (77)

Then Eq. (74) can be numerically solved for the rest trajectories of null geodesics. After getting the trajectories, the lengths of trajectories can be obtained by integrating the line element

dl=1μ1μ2(dμdφ)2+1dφ.\displaystyle dl=\frac{1}{\mu}\sqrt{\frac{1}{\mu^{2}}\left(\frac{d\mu}{d\varphi}\right)^{2}+1}d\varphi\,. (78)

Then the asymptotic time of the light rays can be further obtained by

t=dlc,\displaystyle t=\int\frac{dl}{c}\,, (79)

where cc is the wave speed observed by asymptotic observer c=dl/dt=1/(14ψ)c=dl/dt=\sqrt{1/(1-4\psi)}.

However, in practice, Eq. (75) has to be set at a finite radius, where Eq. (77) should be small enough and its impact on the incident rays can be neglected. For this purposes, in Figure 10 we show the right-hand side of Eq.( 74) as a function of radius rr. In this calculation, we take M=105MM=10^{5}M_{\odot}, the same as in the main text. For r2Mr\gg 2M, μ[μ2ζ(M,μ)]\frac{\partial}{\partial\mu}\left[\mu^{2}\zeta(M,\mu)\right] decays as 1/r21/r^{2}, which is faster than that of the potential. As a result, when r>10[sec]r>10[\rm sec], μ[μ2ζ(M,μ)]\frac{\partial}{\partial\mu}\left[\mu^{2}\zeta(M,\mu)\right] becomes negligible.

Figure 11 shows the trajectories of null geodesics in the spacetime of the scatterer. The initial conditions of Eq. (75) are set at z=40[sec]z=-40[\rm sec], 50[sec]50[\rm sec] away from the scatterer. The red curves indicate the wavefront at different killing times tt. Since Eq. (74) can not be applied to the zz-axis, we do not show light rays on the zz-axis. From Figure 11, the impact of the scatterer on wavefront is only significant when the wavefront is close to the scatter within a radius of 10[sec]10[\rm sec]. Outside this region, the distortion effect on the wavefront is negligible.

Appendix C Analytical solutions for wave equations

We work with the paraboloidal coordinates

{ξ=rz,η=r+z.\left\{\begin{aligned} \xi&=r-z\,,\\ \eta&=r+z\,.\end{aligned}\right. (80)

For ψ=M/r\psi=-M/r, Eq. (47) can be cast into

4ξ+η[ξ(ξhξ)+η(ηhη)+1ξη2hϕ2]+ω2h+4ω2Mξ+ηh=0.\displaystyle\frac{4}{\xi+\eta}\left[\frac{\partial}{\partial\xi}\left(\xi\frac{\partial h}{\partial\xi}\right)+\frac{\partial}{\partial\eta}\left(\eta\frac{\partial h}{\partial\eta}\right)+\frac{1}{\xi\eta}\frac{\partial^{2}h}{\partial\phi^{2}}\right]+\omega^{2}h+\frac{4\omega^{2}M}{\xi+\eta}h=0\,. (81)

Since the incident waves are of azimuthal symmetry, hh does not depend on ϕ\phi and is only a function of η\eta and ξ\xi. We, theorfore, choose the following ansatz

h~=eik~zf(ξ)=eik~ηξ2f(ξ).\displaystyle\tilde{h}=e^{i\tilde{k}z}f(\xi)=e^{i\tilde{k}\frac{\eta-\xi}{2}}f(\xi)\,. (82)

Since f(ξ)f(\xi) does not depend on the direction of k~\tilde{k}, h~\tilde{h} only has Fourier modes with wave vectors k~\tilde{k} along the zz-axis. As such, the above ansatz implicitly assumes that the wave vector k~\tilde{k} of the outgoing waves at infinity is still along the zz-axis. Inserting the above expression into Eq. (81), we obtain an equation for ff

ξd2dξ2f+(1ik~ξ)ddξf+Mω2f=0.\displaystyle\xi\frac{d^{2}}{d\xi^{2}}f+(1-i\tilde{k}\xi)\frac{d}{d\xi}f+M\omega^{2}f=0\,. (83)

The above equation has solutions

f=CF11(iωM,1,iωξ),forξ0,\displaystyle f=C{{}_{1}F_{1}}(i\omega M,1,i\omega\xi),\quad for\quad\xi\neq 0\,, (84)

where F11{{}_{1}F_{1}} is known as the confluent hypergeometric function and CC is a constant, which can be determined by boundary conditions. When M=0M=0

F11(0,1,iωξ)=1.\displaystyle{{}_{1}F_{1}}(0,1,i\omega\xi)=1\,.

Wave function h~\tilde{h} simply goes back to the plane wave. Moreover, it is worth noting that F11{{}_{1}F_{1}} is not well defined on the zz-axis (ξ=0\xi=0).

If we focus on regions that far away from the zz-axis where ξ1\xi\gg 1, the confluent hypergeometric F11(iωM,1,iωξ){{}_{1}F_{1}}(i\omega M,1,i\omega\xi) can be expanded around iωξ=i\omega\xi=\infty as

F11(iωM,1,iωξ)ξiωM[1+ω2M2ξ+O(1ξ)2].\displaystyle{{}_{1}F_{1}}(i\omega M,1,i\omega\xi)\approx\xi^{-i\omega M}\left[1+\frac{\omega^{2}M^{2}}{\xi}+O\left(\frac{1}{\xi}\right)^{2}\right]\,. (85)

The above equation can be evaluated efficiently in the numerical processes.