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

thanks: Corresponding author

More Nonlinearities?
II. A Short Guide to First- and Second-Order Electromagnetic Perturbations in the Schwarzschild background.

Fawzi Aly1 [Uncaptioned image] mabbasal[AT]buffalo.edu    Dejan Stojkovic1 [Uncaptioned image] ds77[AT]buffalo.edu 1HEPCOS, Physics Department, SUNY at Buffalo, Buffalo, New York, USA
Abstract

We study second-order electromagnetic perturbations in the Schwarzschild background and derive the effective source terms for Regge-Wheeler equation which are quadratic in first-order gravitational and electromagnetic perturbations. In addition to the induced mixed quadratic modes, we find that linear gravitational modes are also excited, with amplitudes dependent on the electromagnetic potential. A toy model involving a Dirac delta function potential demonstrates mixing of linear gravitational and electromagnetic perturbations with frequencies ω(1)\omega^{(1)} and Ω(1)\Omega^{(1)}, resulting in the second-order QNM mixing in the electromagnetic field at Ω(2)=Ω(1)+ω(1)\Omega^{(2)}=\Omega^{(1)}+\omega^{(1)}. This complements prior work in [1] on the second-order gravitational perturbation mixing and highlights potential applications in multi-messenger astrophysics for systems observed by LIGO and upcoming LISA. We also study first-order perturbations due to a point charge and show it could be reduced to a one-dimensional path integral. Within the toy model, we investigate the first-order electromagnetic perturbation due to a radially free-falling single charge qq and radial dipole moment p=qηp=q\eta, employing semi-analytical and numerical methods. For the dipole case, we show that the QNM perturbation is excited with a nearly constant amplitude. Future work will focus on incorporating mixing in more realistic potentials and exploring numerical approach in the context of rotating spacetimes.

Introduction

The study of quasinormal modes (QNMs) is fundamental to understanding how black holes (BHs) respond to perturbations. In general relativity (GR), QNMs play an even more critical role, as they are only parametrized by BH macroscopic properties: mass MM, angular momentum JJ, and charge QQ—making them characteristic signatures of these enigmatic objects [2, 3, 4, 5]. In this context, the black hole spectroscopy program was established. While observing a single QNM provides a direct measurement of these parameters, detecting two or more QNMs offers a valuable consistency check. This enables testing the validity limits of black hole perturbation theory (BHPT) in modeling the late-stage ringdown phase [2, 6, 7].

While gravitational QNMs have garnered significant attention—particularly following the groundbreaking gravitational wave (GW) detections by LIGO and Virgo [8, 9]—electromagnetic QNMs remain a relatively under-explored area with intriguing theoretical implications, and potential astrophysical relevance. This is particularly true in studying neutron stars (NSs), and charged BH properties. Electromagnetic QNMs encode some information about how electromagnetic fields evolve and decay around vacuum BH spacetimes, for example like those considered as perturbations of Kerr type remnants after magnetized mergers [10]. Understanding these modes is essential for better interpretation of electromagnetic radiation spectrum emitted during astrophysical events, including black hole-neutron star (NSBH) and binary neutron star (BNS) mergers, gravitational collapse of unstable magnetars, magnetar flares, and accretion processes near BHs.

In the context of multi-messenger astronomy [11, 12, 13, 14], the simultaneous detection of gravitational and electromagnetic signals offers a unique opportunity to probe the strong-field regime and more thoroughly examine astrophysical events where electromagnetic effects are switched on. The gravitational and electromagnetic signals are expected to provide both complementary and overlapping information about the same astrophysical event, allowing for cross-checks. Although, in principle, understanding electromagnetic QNM spectra might lead to more accurate extraction of information in the aforementioned systems [15, 5] and enable further testing of GR and its coupling with electromagnetism [1], the state of the art in any realistic signal processing for BNS is to ignore the QNM-component. This is partially due to the complexity of magnetized mergers and lack of full description of how electromagnetic field behaves in the vicinity of the ejecta and in plasma environment.

In the context of numerical simulations, scenarios involving mergers of charged binary BHs or BNS, magnetar collapse, and NSBH systems without significant disruption offer a higher probability of resolving electromagnetic QNM behavior. Current and future simulations of charged or magnetized systems could shed light on the remnant’s ringdown for these scenarios, especially following recent attempts to identify electromagnetic QNMs, (i.e. in [16, 17]). Given that magnetic fields in astrophysical BNS mergers can reach up to B1012TB\sim 10^{12}\,\text{T} [18], it is intriguing to explore the role of electromagnetic fields in these extreme conditions.

In [1], we explored whether electromagnetic fields can leave imprints on gravitational radiation, focusing more on extreme mass ratio inspiral (EMRI) scenarios [19, 20, 21, 22]. We reported that it was plausible for electromagnetic effects to compete with the second-order gravitational perturbations in a NSBH merger. With a lower limit estimate for astrophysical NS of mass 2M2M_{\odot}, and with a magnetic field of order 107T\sim 10^{7}\,\text{T}, in an inspiral around a supermassive black hole of mass 106M10^{6}M_{\odot}. However, in the context of stellar-mass mergers, the situation is more challenging in the absence of a proper way to define perturbation parameters, so perhaps further analytical and numerical investigations are needed.

However, since gravitational effects are expected to dominate in astrophysics, this complementary work aims to generically investigate the reverse question: Can we detect gravitational signatures in the electromagnetic radiation? Such phenomena should be important in analyzing and modeling electromagnetic counterparts of magnetized collapse or mergers in future.

The work in this paper addresses several questions. First, we provide a brief review of the Regge-Wheeler formalism for electromagnetic perturbations. We then study second-order electromagnetic perturbations, where the electromagnetic-gravitational interaction, with its first-order modes, acts as an effective source term. We solve this problem for a sample source term for the purpose of illustration. Following this, we go back to first order perturbation, and solve for the electromagnetic perturbations induced by a point charge and also an ideal dipole, after approximating the Regge-Wheeler potential with a Dirac delta function. The choice of an ideal dipole serves as an analogy for astrophysical systems with relatively small macroscopic charges. Finally, we report our numerical solutions for the perturbation due to the dipole.

The structure of this paper is as follows. In Section I, we outline the electromagnetic perturbation in Schwarzschild coordinates. Next, in Section II, we formulate the second-order perturbation and derive the source terms for the second-order master Regge-Wheeler formalism for spin s=1s=1. By the end of this section, we solve for a sample source using the Dirac delta function approximation to illustrate electromagnetic-gravitational mixing, similar to [1]. Then, we return to first-order perturbations in Section III and show that, in the time domain, the inhomogeneous solution due to a point charge can be reduced to a one-dimensional path integral, and we discuss several examples. In Sections IV and V, we investigate the first-order electromagnetic perturbation due to a radially free-falling point charge and then a radial dipole in the Dirac delta approximation using a semi-analytical approach. Next, we also solve for the perturbation due to the dipole numerically and comment on the solution in Section VI. Finally, we discuss the results of our analysis and potential directions for future work in Section VII. Throughout this work, we use natural units where G=c=1G=c=1.

I Preliminary

In this section, we briefly review the spherical decomposition of Maxwell’s equations on a spherically symmetric spacetime, such as Schwarzschild. We discuss how the Faraday tensor FF can be expressed in terms of perturbation quantities Φlm\Phi_{lm}, as well as the current multipoles jlmj_{lm}, and then obtain the corresponding electromagnetic stress-energy tensor TT. These few steps will allow us to solve the toy model introduced in IV.

I.1 Maxwell equations

To derive the Regge-Wheeler equations for electromagnetic perturbations in a Schwarzschild spacetime [10, 23, 24, 25], the vector potential AμA_{\mu} and the current vector JμJ_{\mu} are decomposed into even and odd components, each of which is expanded in terms of 4-vector spherical harmonics, as outlined in Appendix D. Consequently, the Faraday tensor FμνF_{\mu\nu}, and therefore Maxwell’s equations, are similarly divided into sectors of even and odd parity. The first order even and odd electromagnetic perturbations

Φlm(1)e(t,x)=r2λ(hlmtglmr)\displaystyle{}_{e}\Phi^{(1)}_{lm}(t,x)=\frac{r^{2}}{\lambda}\left(\frac{\partial h_{lm}}{\partial t}-\frac{\partial g_{lm}}{\partial r}\right) (1)
Φlm(1)o(t,x)=alm\displaystyle{}_{o}\Phi^{(1)}_{lm}(t,x)=a_{lm}

are governed by the same potential VRWs=1=λf(r)r2{}_{s=1}V_{RW}=\frac{\lambda f(r)}{r^{2}}, where λ=l(l+1)\lambda=l(l+1) and ll represents the angular number, but differ in how the source term 𝒮o,elm(1)J(t,x){}_{J}\mathcal{S}^{(1)}_{o,e\,lm}(t,x) is derived from JμJ_{\mu}. The governing equation for these perturbations is expressed as:

EM:=xxttVRWs=1(x),\displaystyle\mathcal{L}_{EM}:=\partial_{xx}-\partial_{tt}-\,{}_{s=1}V_{RW}(x), (2)
EMΦlm(1)o,e(t,x)=𝒮o,elm(1)J(t,x),\displaystyle\mathcal{L}_{EM}\,{}_{o,e}\Phi^{(1)}_{lm}(t,x)={}_{J}\mathcal{S}^{(1)}_{o,e\,lm}(t,x),

where xx is the tortoise radial coordinate, related to the areal radial coordinate through dr=f(r)dxdr=f(r)dx, while tt is the time measured by a clock asymptotically far from the Schwarzschild black hole. In addition, 𝒮o,elm(1)J(t,x){}_{J}\mathcal{S}^{(1)}_{o,e\,lm}(t,x) is formulated in terms of the current vector multipoles j(i)lmj_{(i)\,lm},where i=1,2,3,4i=1,2,3,4, with an example provided in section III.

This ODE governing the perturbations has two regular singularities (rank = 1) at r=0r=0 and r=2Mr=2M, along with an irregular singularity (rank = 2) as rr\rightarrow\infty. This structure classifies it as a confluent Heun ODE [26, 19], which can be solved using confluent Heun functions HCH_{C} or, in some cases, a narrower subset of confluent Heun polynomials [27]. The QNMs boundary conditions are imposed on the solution, and together with two initial conditions, the full solution can be specified, the reader can find all the details in Section II.A of [1].

I.2 Faraday tensor

After successfully separating and decoupling the system into its two degrees of freedom Φlm(1)o,e{}_{o,e}\Phi^{(1)}_{lm}, the Faraday tensor FμνF_{\mu\nu} can now be expressed in terms Φlm(1)o,e{}_{o,e}\Phi^{(1)}_{lm} and the current multipoles j(i)lm(t,x)j_{(i)\,lm}(t,x) for l>0l>0. However, before proceeding, let’s first address the monopole term l=0l=0 of FμνF_{\mu\nu}.

I.2.1 Monopole Case

Given the multipole decomposition, the monopole case can be solved separately, as the potential vanishes. The coupled equations for the scalar field Φ00\Phi_{00} are:

Φ00r\displaystyle\frac{\partial\Phi_{00}}{\partial r} =r2j(1) 00f(r)\displaystyle=\frac{r^{2}\,j_{(1)\,00}}{f(r)} (3)
Φ00t\displaystyle\frac{\partial\Phi_{00}}{\partial t} =r2f(r)j(2) 00.\displaystyle=r^{2}f(r)\,j_{(2)\,00}.

When the perturbation is due to a collection of discrete charges, the solution reduces to a summation over the individual charges qiq_{i}, each multiplied by a step function dependent on their rrpi(t)r-r_{p_{i}}(t) argument

Φ00=iqiΘ[rrpi(t)],\Phi_{00}=\sum_{i}q_{i}\,\Theta[r-r_{p_{i}}(t)], (4)

where Θ\Theta denotes the Heaviside step function and rpi(t)r_{p_{i}}(t) is the radial coordinate of the ii-th charge parameterized by tt. This solution should resemble a static perturbation in the background spacetime at radii greater than the position of the charge distribution. Such a perturbation effectively charges the spacetime, transforming the region outside the charge distribution into a member of the Reissner–Nordström family of spacetimes[28, 29, 30]:

Fμν00=(0Qtot(t,r)r20000000000000)F_{\mu\nu 00}=\begin{pmatrix}0&\frac{Q_{\text{tot}}(t,r)}{r^{2}}&0&0\\ **&0&0&0\\ 0&0&0&0\\ 0&0&0&0\end{pmatrix} (5)

where Qtot=iqiΘ[rrpi(t)]Q_{\text{tot}}=\sum_{i}q_{i}\,\Theta[r-r_{p_{i}}(t)], somewhat analogous to Gauss’s law in electrostatics in the exterior region. It is important to note that if the source includes no monopole contribution, this perturbation vanishes, as in the case of dipole moments when modeling astrophysical systems with relatively small macroscopic charge relative to the higher moments. Subsequently, any electromagnetic moments would be radiated away on a time scale comparable to that of the merger and post-merger phases 111We are assuming that the remnant spacetime exterior is vacuum. However, if there were accretion disks or plasma present, the situation might differ, and the same conclusions may not hold.. During the process when the BHPT is expected to hold, the Φlm(1)o,e{}_{o,e}\Phi^{(1)}_{lm} should describe this radiation. Furthermore, if the electromagnetic moments are sufficiently large, this radiation could be substantial, albeit subdominant to gravitational radiation. In such a case, coupling effects between the two radiations must be considered, which is the purpose of this work and its companion [1].

I.2.2 Multipoles l>0l>0

For higher multipoles l>0l>0, the odd components of the Maxwell tensor Fμνlmo{}_{o}F_{\mu\nu lm} in the multipole expansion take the following forms for a generic source term:

Flmo=(00cscθYlmΦlmotcscθYlmΦlmor00sinθY˙lmΦlmotsinθY˙lmΦlmor0λsinθYlmΦlmo0).{}_{o}F_{lm}=\begin{pmatrix}0&0&-\csc\theta Y_{lm}^{\prime}\frac{\partial{}_{o}\Phi_{lm}}{\partial t}&-\csc\theta Y_{lm}^{\prime}\frac{\partial{}_{o}\Phi_{lm}}{\partial r}\\ 0&0&\sin\theta\dot{Y}_{lm}\frac{\partial{}_{o}\Phi_{lm}}{\partial t}&\sin\theta\dot{Y}_{lm}\frac{\partial{}_{o}\Phi_{lm}}{\partial r}\\ **&**&0&-\lambda\,\sin\theta Y_{lm}\,{}_{o}\Phi_{lm}\\ **&**&**&0\end{pmatrix}. (6)

The even part of the Maxwell tensor could be divided into two contributions, which can be expressed as:

Fμνlme=Fμνlme1+Fμνlme2,{}_{e}F_{\mu\nu lm}={}_{e1}F_{\mu\nu lm}+{}_{e2}F_{\mu\nu lm},

where the first contribution, Fμνlme1{}_{e1}F_{\mu\nu lm}, is written in terms of the even scalar perturbation Φe{}_{e}\Phi, while the second contribution, Fμνlme2{}_{e2}F_{\mu\nu lm}, is expressed in terms of the j(1)lmj_{(1)\,lm} and j(2)lmj_{(2)\,lm} multipole components of the even current vector.

Flme1=(0λr2YlmΦlmeY˙lmfΦlmerYlmfΦlmer0Y˙lmΦlmet1fYlmΦlmet1f000),{}_{e1}F_{lm}=\begin{pmatrix}0&\frac{\lambda}{r^{2}}Y_{lm}{}_{e}\Phi_{lm}&\dot{Y}_{lm}f\frac{\partial{}_{e}\Phi_{lm}}{\partial r}&Y^{\prime}_{lm}f\frac{\partial{}_{e}\Phi_{lm}}{\partial r}\\ **&0&\dot{Y}_{lm}\frac{\partial{}_{e}\Phi_{lm}}{\partial t}\frac{1}{f}&Y^{\prime}_{lm}\frac{\partial{}_{e}\Phi_{lm}}{\partial t}\frac{1}{f}\\ **&**&0&0\\ **&**&**&0\end{pmatrix}, (7)

and,

Flme2=(00Y˙lmr2j(1)lmλYlmr2j(1)lmλ0Y˙lmr2j(2)lmλYlmr2j(2)lmλ000).{}_{e2}F_{lm}=\begin{pmatrix}0&0&-\dot{Y}_{lm}\frac{r^{2}j_{(1)lm}}{\lambda}&-Y^{\prime}_{lm}\frac{r^{2}j_{(1)lm}}{\lambda}\\ **&0&-\dot{Y}_{lm}\frac{r^{2}j_{(2)lm}}{\lambda}&-Y^{\prime}_{lm}\frac{r^{2}j_{(2)lm}}{\lambda}\\ **&**&0&0\\ **&**&**&0\end{pmatrix}. (8)

Finally we can express the Maxwell tensor FμνF_{\mu\nu} as

Fμν\displaystyle F_{\mu\nu} =Fμν 00+l=1m=llFμνlme2\displaystyle=F_{\mu\nu\,00}+\sum_{l=1}^{\infty}\sum_{m=-l}^{l}{}_{e2}F_{\mu\nu\,lm} (9)
+l=1m=ll(Fμνlme1+Fμνlmo)\displaystyle+\sum_{l=1}^{\infty}\sum_{m=-l}^{l}\left({}_{e1}F_{\mu\nu\,lm}+{}_{o}F_{\mu\nu\,lm}\right)

So far, we haven’t made any specific assumptions about the perturbation source, yet we can make some general statements about the behavior of the Maxwell tensor in its current form.

  • While solving the system (3), it becomes clear that the monopole Φ00\Phi_{00} neither exhibits radiative behavior nor satisfies the typical boundary conditions imposed on QNMs. As tt\rightarrow\infty on the clock of an asymptotic observer, the charge will freeze near the horizon, all radiative components (QNMs,tails,..etc) will dissipate. Ultimately, the system will asymptotically approach the Reissner-Nordström solution.

  • The term Fμνlme2{}_{e2}F_{\mu\nu\,lm} contains more information about the behavior of the electromagnetic field in the neighborhood of the charge, typically localized if the source JμJ_{\mu} behaves as expected in astrophysical cases. Consequently, when calculating the stress-energy tensor TμνT_{\mu\nu}, this component needs to be regularized for discrete charge distributions to avoid infinite energy density. We shall refer to Fμνlme2{}_{e2}F_{\mu\nu\,lm} as “internal” field from now on.

  • The components Fμνlme1{}_{e1}F_{\mu\nu\,lm} and Fμνlmo{}_{o}F_{\mu\nu\,lm} are radiative in nature and represent the QNM behavior of the electromagnetic field. These components are obtained by solving the inhomogeneous Regge-Wheeler equations (2), where j(1)lmj_{(1)\,lm} and j(2)lmj_{(2)\,lm} source the even component, and j(4)lmj_{(4)\,lm} sources the odd component. The initial conditions on the Electromagnetic field will add non-trivial constraints on Φlm(1)o,e{}_{o,e}\Phi^{(1)}_{lm}, this will be highly relevant if we are focusing more on stellar-mass mergers (i.e. NSBH or BNS mergers) ringdown phase.

I.3 Stress-energy tensor

Given the Faraday tensor FμνF_{\mu\nu}, we can proceed to calculate the electromagnetic stress-energy tensor TEMμνT_{EM}^{\mu\nu}, which describes the distribution and flow of electromagnetic energy and momentum in spacetime. In spherical coordinates, TEMμνT_{EM}^{\mu\nu} takes the following standard form:

TEMμν=(ϵSrSθSϕSrPrσrθσrϕSθσrθPθσθϕSϕσrϕσθϕPϕ,)T_{EM}^{\mu\nu}=\begin{pmatrix}\epsilon&S_{r}&S_{\theta}&S_{\phi}\\ S_{r}&P_{r}&\sigma_{r\theta}&\sigma_{r\phi}\\ S_{\theta}&\sigma_{r\theta}&P_{\theta}&\sigma_{\theta\phi}\\ S_{\phi}&\sigma_{r\phi}&\sigma_{\theta\phi}&P_{\phi},\\ \end{pmatrix} (10)

Assuming that the components of the stress-energy tensor retain their familiar physical interpretations as they do in Minkowski spacetime, ϵ\epsilon represents the energy density or electromagnetic energy per unit volume; SrS_{r}, SθS_{\theta}, and SϕS_{\phi} are the components of the Poynting vector, indicating the flow of electromagnetic energy in the radial, polar, and azimuthal directions, respectively; PrP_{r}, PθP_{\theta}, and PϕP_{\phi} correspond to the pressure in these directions; and the off-diagonal elements σrθ\sigma_{r\theta}, σrϕ\sigma_{r\phi}, and σθϕ\sigma_{\theta\phi} describe the shear stresses in the electromagnetic field.

As discussed earlier, we will exclude the contribution from Fμνlme1{}_{e1}F_{\mu\nu\,lm} when calculating TEMμνT_{EM}^{\mu\nu}. Since the full expressions are lengthy, for completeness and illustration purposes, we will just write down the expressions for ϵ\epsilon and SrS_{r} while rest will be provided in a companion Mathematica Notebook:

ϵ\displaystyle\epsilon =l,ml,mAlmlm2r2f(r)[ΦlmeΦlmorΦlmerΦlmo\displaystyle=\sum_{l,m}\sum_{l^{\prime},m^{\prime}}\frac{A_{lml^{\prime}m^{\prime}}}{2r^{2}f(r)}\bigg{[}{}_{e}\Phi_{lm}\frac{\partial{}_{o}\Phi_{l^{\prime}m^{\prime}}}{\partial r}-\frac{\partial{}_{e}\Phi_{lm}}{\partial r}{}_{o}\Phi_{l^{\prime}m^{\prime}} (11)
+ΦlmoΦlmer+ΦlmorΦlme]\displaystyle\quad+{}_{o}\Phi_{lm}\frac{\partial{}_{e}\Phi_{l^{\prime}m^{\prime}}}{\partial r}+\frac{\partial{}_{o}\Phi_{lm}}{\partial r}{}_{e}\Phi_{l^{\prime}m^{\prime}}\bigg{]}
+Blmlm2r2[ΦlmorΦlmorΦlmerΦlmer\displaystyle+\frac{B_{lml^{\prime}m^{\prime}}}{2r^{2}}\bigg{[}-\frac{\partial{}_{o}\Phi_{lm}}{\partial r}\frac{\partial{}_{o}\Phi_{l^{\prime}m^{\prime}}}{\partial r}-\frac{\partial{}_{e}\Phi_{lm}}{\partial r}\frac{\partial{}_{e}\Phi_{l^{\prime}m^{\prime}}}{\partial r}
1f(r)2ΦlmotΦlmot1f(r)2ΦlmetΦlmet]\displaystyle\quad-\frac{1}{f(r)^{2}}\frac{\partial{}_{o}\Phi_{lm}}{\partial t}\frac{\partial{}_{o}\Phi_{l^{\prime}m^{\prime}}}{\partial t}-\frac{1}{f(r)^{2}}\frac{\partial{}_{e}\Phi_{lm}}{\partial t}\frac{\partial{}_{e}\Phi_{l^{\prime}m^{\prime}}}{\partial t}\bigg{]}
λ2YlmYlm(ΦlmeΦlme+ΦlmoΦlmo)2r4f(r)\displaystyle-\frac{\lambda^{2}Y_{lm}Y_{l^{\prime}m^{\prime}}\left({}_{e}\Phi_{lm}{}_{e}\Phi_{l^{\prime}m^{\prime}}+{}_{o}\Phi_{lm}{}_{o}\Phi_{l^{\prime}m^{\prime}}\right)}{2r^{4}f(r)}
Qtotr4f(r)l,mλYlmΦlmeQtot22r4f(r)\displaystyle-\frac{Q_{tot}}{{r^{4}f(r)}}\sum_{l,m}\lambda Y_{lm}{}_{e}\Phi_{lm}-\frac{Q_{tot}^{2}}{2r^{4}f(r)}

and

Sr\displaystyle S_{r} =1r2l,ml,m{Almlm[f(r)ΦlmerΦlmor\displaystyle=\frac{1}{r^{2}}\sum_{l,m}\sum_{l^{\prime},m^{\prime}}\bigg{\{}A_{lml^{\prime}m^{\prime}}\left[f(r)\frac{\partial{}_{e}\Phi_{lm}}{\partial r}\frac{\partial{}_{o}\Phi_{l^{\prime}m^{\prime}}}{\partial r}\right. (12)
1f(r)ΦlmotΦlmet]+Blmlm[ΦlmorΦlmot\displaystyle\left.-\frac{1}{f(r)}\frac{\partial{}_{o}\Phi_{lm}}{\partial t}\frac{\partial{}_{e}\Phi_{l^{\prime}m^{\prime}}}{\partial t}\right]+B_{lml^{\prime}m^{\prime}}\left[\frac{\partial{}_{o}\Phi_{l^{\prime}m^{\prime}}}{\partial r}\frac{\partial{}_{o}\Phi_{lm}}{\partial t}\right.
+ΦlmetΦlmer]}\displaystyle\left.+\frac{\partial{}_{e}\Phi_{l^{\prime}m^{\prime}}}{\partial t}\frac{\partial{}_{e}\Phi_{lm}}{\partial r}\right]\bigg{\}}

where

Almlm=cscθ[Y˙lmYlmYlmY˙lm],Blmlm=csc2θ[YlmYlm+Y˙lmY˙lm]\begin{gathered}A_{lml^{\prime}m^{\prime}}=\csc\theta\left[\dot{Y}_{lm}Y^{\prime}_{l^{\prime}m^{\prime}}-Y^{\prime}_{lm}\dot{Y}_{l^{\prime}m^{\prime}}\right],\\ B_{lml^{\prime}m^{\prime}}=\csc^{2}\theta\left[Y^{\prime}_{lm}Y^{\prime}_{l^{\prime}m^{\prime}}+\dot{Y}_{lm}\dot{Y}_{l^{\prime}m^{\prime}}\right]\end{gathered} (13)

The main takeaways from this brief summary so far are as follows:

  • The stress-energy tensor TEMμνT_{EM}^{\mu\nu} involves both linear terms in Φlme{}_{e}\Phi_{lm} and Φlmo{}_{o}\Phi_{lm}, as well as quadratic terms involving products of the form Φlme×Φlme{}_{e}\Phi_{lm}\times{}_{e}\Phi_{l^{\prime}m^{\prime}}, Φlmo×Φlmo{}_{o}\Phi_{lm}\times{}_{o}\Phi_{l^{\prime}m^{\prime}}, and Φlme×Φlmo{}_{e}\Phi_{lm}\times{}_{o}\Phi_{l^{\prime}m^{\prime}} and their derivatives, summed over the indices ll, ll^{\prime}, mm, and mm^{\prime}. As a result, we can anticipate that the frequencies associated with the mixing modes will be either linear or quadratic as argued for in [1].

  • We can decompose equation (10) using, for example, the 4D tensor harmonics defined as in [1]. In the Laplace spectral space (or less formally in the Fourier space), this will allow us to study the spectral decomposition of the energy flux, shear, and density as functions of rr for each multipole lmlm and frequency Ωn\Omega_{n} similar to [31, 32]. Furthermore, we can examine how they couple to other radiative multipoles lml^{\prime}m^{\prime} of Φlm\Phi_{l^{\prime}m^{\prime}}.

  • The spherical decomposition of TEMμνT_{EM}^{\mu\nu} in expression (9) is not fully compatible with gravitational perturbations. To resolve this, we propose projecting it onto the tensorial spherical harmonics provided in Appendix A of [1] or following ref [23] as demonstrated in Appendix C,D,E, and F in [1], ensuring compatibility with the Regge-Wheeler gravitational analysis[1].

  • As tt\rightarrow\infty, both Φ\Phi and its derivatives will vanish, and the total charge QtotQ_{\text{tot}} will appear frozen near the horizon from the perspective of an asymptotic observer. Consequently, only the monopole term, i.e., the Reissner–Nordström charge, will be measurable at “late times”. However, as different radiative modes (i.e. QNMs and tails) have different decaying rates, that suggest the we should be careful what time-scale we are interested in studying.

Refer to caption
Figure 1: Comparison of gravitational and electromagnetic QNMs in Schwarzschild spacetime. Gravitational frequencies ωln(1)\omega^{(1)}_{ln} for l=2,3l=2,3 and n=0,1n=0,1 are compared with electromagnetic frequencies Ωln(1)\Omega^{(1)}_{ln} for l=1,2l=1,2 and n=0,1n=0,1. For the quadratic QNMs, the real part of the frequencies is given by the sum Ω(i×j)R(2)=Ω(i)R(1)±ω(j)R(1)\Omega^{(2)}_{(i\times j)R}=\Omega^{(1)}_{(i)R}\pm\omega^{(1)}_{(j)R} of the linear modes, while the imaginary part, representing the reciprocal of the decay time, is given by Ω(i×j)I(2)=Ω(i)I(1)+ω(j)I(1)\Omega^{(2)}_{(i\times j)I}=\Omega^{(1)}_{(i)I}+\omega^{(1)}_{(j)I} as a sum of the corresponding linear terms. The blue squares and red dots represent the electromagnetic and gravitational QNMs, respectively, while cyan rhombuses indicate the quadratic QNMs.

II Second order electromagnetic perturbation

In [1], the coupling between electromagnetic and gravitational modes was investigated perturbatively in the framework of minimally coupling Einstein-Maxwell system. It was reported that, in the simplest case of mixing between the two fields, when the electromagnetic field sources gravitational perturbations, it excites gravitational modes that are linear and quadratic in the electromagnetic QNMs. Also amplitude of the linear gravitational QNMs get correction from electromagnetic field. As we go higher in mixing, more coupling between the modes is expected to take place, such that modes quadratic in both fields arise.

On the other hand, at first and second order in α\alpha, the Maxwell equation for the perturbed electromagnetic field is expressed as:

αμ(0)F(1)μν=αJ(1)να2μ(0)F(2)μν=α2Jeff(2)ν,\begin{gathered}\alpha\nabla_{\mu}^{(0)}F^{(1)\mu\nu}=\alpha J^{(1)\nu}\\ \alpha^{2}\nabla_{\mu}^{(0)}F^{(2)\mu\nu}=\alpha^{2}J_{eff}^{(2)\,\nu},\end{gathered} (14)

where Fμν(1,2)F^{(1,2)}_{\mu\nu} and J(1,2)νJ^{(1,2)\nu} represent the first and second -order perturbation of the electromagnetic field and second-order current source term respectively. The covariant derivative μ(0)\nabla_{\mu}^{(0)} is taken with respect to the unperturbed Schwarzschild metric gμν(0)g_{\mu\nu}^{(0)}, hence it’s straightforward to reduce the second order field perturbation to their corresponding Regge-Wheeler second odd Φlm(2)o{}_{o}\Phi^{(2)}_{lm} and even Φlm(2)e{}_{e}\Phi^{(2)}_{lm} perturbation:

EMΦlm(2)o,e(t,x)=𝒮o,elm(2)eff(t,x).\mathcal{L}_{EM}\,{}_{o,e}\Phi^{(2)}_{lm}(t,x)={}_{\text{eff}}\mathcal{S}^{(2)}_{o,e\,lm}(t,x).

At the second order, there is additional effective nonlinear contribution which Jeff(2)νJ_{\text{eff}}^{(2)\nu} represents coupling between the first-order electromagnetic field Fμν(1)F^{(1)}_{\mu\nu} and the first-order metric perturbation gμν(1)g^{(1)}_{\mu\nu}. The effective second-order current, Jeff(2)νJ_{\text{eff}}^{(2)\nu}, is given by

Jeff(2)ν\displaystyle J_{\text{eff}}^{(2)\nu} =12(g(0)γμF(1)αν+g(0)νγF(1)μα)\displaystyle=-\frac{1}{2}\left(g^{(0)\gamma\mu}F^{(1)\alpha\nu}+g^{(0)\nu\gamma}F^{(1)\mu\alpha}\right) (15)
×(μ(0)gαγ(1)+α(0)gμγ(1)γ(0)gαμ(1)),\displaystyle\times\left(\nabla_{\mu}^{(0)}g_{\alpha\gamma}^{(1)}+\nabla_{\alpha}^{(0)}g_{\mu\gamma}^{(1)}-\nabla_{\gamma}^{(0)}g_{\alpha\mu}^{(1)}\right),
=12F(1)ναα(0)gμμ(1)\displaystyle=\frac{1}{2}F^{(1)\nu\alpha}\nabla_{\alpha}^{(0)}g_{\mu}^{\mu\,(1)}

It is straightforward to find the corresponding Regge-Wheeler sources corresponding to Jeff(2)νJ_{\text{eff}}^{(2)\nu}, through using vector spherical harmonics in Appendix D. We report the final result for even and odd sources respectively here

Selmeff=12λ(2t22x2)j~lmj~lm=[2(x+A¯)ψelm][Qtotδllδmm+(1)mλ′′Φel′′m′′Bl,m;l,m,l′′,m′′],A¯=2r(λ2)+6M[(12λ1)12λ+3Mr(12λf)].\begin{gathered}{}_{\text{eff}}S_{e\,lm}=\frac{1}{2\lambda}\left(\frac{\partial^{2}}{\partial t^{2}}-\frac{\partial^{2}}{\partial x^{2}}\right)\tilde{j}_{lm}\\ \tilde{j}_{lm}=\left[-2\left(\frac{\partial}{\partial x}+\bar{A}\right)\psi_{e\,l^{\prime}m^{\prime}}\right]\left[Q_{\text{tot}}\,\delta_{ll^{\prime}}\delta_{mm^{\prime}}+(-1)^{m}\,\lambda^{\prime\prime}\,\Phi_{e\,l^{\prime\prime}m^{\prime\prime}}\,B_{l,-m;l^{\prime},m^{\prime},l^{\prime\prime},m^{\prime\prime}}\right],\\ \bar{A}=\frac{2}{r(\lambda-2)+6M}\left[\left(\frac{1}{2}\lambda-1\right)\frac{1}{2}\lambda+\frac{3M}{r}\left(\frac{1}{2}\lambda-f\right)\right].\end{gathered} (16)
Solmeff=(1)mf2{Cl,m;l,m,l′′,m′′λ′′Φol′′m′′r2[2(x+A¯)ψelm],+gaaxa[2(x+A¯)ψelm][Cl,m;l′′,m′′,l,mΦol′′m′′xa+Dl,m;l′′,m′′,l,m,(ϵabΦel′′m′′xb+r2j(a)l′′m′′λ′′)]}ϵab=(0f1f0).\begin{gathered}{}_{\text{eff}}S_{o\,lm}=(-1)^{m}\frac{f}{2}\Bigg{\{}-C_{l,-m;l^{\prime},m^{\prime},l^{\prime\prime},m^{\prime\prime}}\frac{\lambda^{\prime\prime}\Phi_{o\,l^{\prime\prime}m^{\prime\prime}}}{r^{2}}\left[-2\left(\frac{\partial}{\partial x}+\bar{A}\right)\psi_{e\,l^{\prime}m^{\prime}}\right],\\ +g^{aa}\frac{\partial}{\partial x_{a}}\left[-2\left(\frac{\partial}{\partial x}+\bar{A}\right)\psi_{e\,l^{\prime}m^{\prime}}\right]\Bigg{[}C_{l,-m;l^{\prime\prime},m^{\prime\prime},l^{\prime},m^{\prime}}\frac{\partial\Phi_{o\,l^{\prime\prime}m^{\prime\prime}}}{\partial x_{a}}+D_{l,-m;l^{\prime\prime},m^{\prime\prime},l^{\prime},m^{\prime}},\left(\epsilon_{a}^{b}\frac{\partial\Phi_{e\,l^{\prime\prime}m^{\prime\prime}}}{\partial x_{b}}+\frac{r^{2}j_{(a)\,l^{\prime\prime}m^{\prime\prime}}}{\lambda^{\prime\prime}}\right)\Bigg{]}\Bigg{\}}\\ \epsilon_{a}^{b}=\begin{pmatrix}0&f\\ \frac{1}{f}&0\end{pmatrix}.\end{gathered} (17)

Here, Bl,m,l,m,l′′,m′′B_{l,m,l^{\prime},m^{\prime},l^{\prime\prime},m^{\prime\prime}}, Cl,m,l,m,l′′,m′′C_{l,m,l^{\prime},m^{\prime},l^{\prime\prime},m^{\prime\prime}}, and Dl,m,l,m,l′′,m′′D_{l,m,l^{\prime},m^{\prime},l^{\prime\prime},m^{\prime\prime}} are the multipolar mixing coefficients provided in Appendix D of [1]. In this context, we have ignored the source term for the linear gravitational perturbation to simplify the expression. Including these terms would contribute a source component independent of Φ\Phi or ψ\psi, as well as the linear electromagnetic modes Φ(1)\Phi^{(1)}, which depend on the specifics of the electromagnetic potential.

Generically, beyond linear electromagnetic perturbations, at higher orders Φ(j),j2\Phi^{(j)},j\geq 2, gravitational perturbations will similarly source higher-order electromagnetic modes. A similar analysis, as outlined in Section II of [1], can be applied here. For instance, at second order in electromagnetic QNMs, we expect to observe phenomena that only arise with higher-level mixing between the two fields, as seen in the gravitational case. For example at second order in α\alpha, if a gravitational linear QNM has a frequency ωi(1)\omega^{(1)}_{i} and a linear electromagnetic QNM has a frequency Ωj(1)\Omega^{(1)}_{j}, part of the quadratic electromagnetic spectrum will include frequencies quadratic in both gravitational-electromagnetic modes Ωi(2)=Ωj(1)±ωk(1)\Omega^{(2)}_{i}=\Omega^{(1)}_{j}\pm\omega^{(1)}_{k} as shown in Fig. 1, however quadratic modes in electromagnetic will be absent.

In most astrophysical scenarios, such as gravitational collapse or mergers, gravitational effects are expected to dominate. In this context, searching for the mixing between electromagnetic and gravitational fields suggests that electromagnetic QNMs with gravitational imprints may be more interesting than the reverse, but the richness of typical BNS and other magnetized phenomena will make it extremely hard to resolve the aforementioned effect. Nevertheless, both approaches could serve as tests for minimal coupling within the framework of General Relativity or for more complex couplings in modified gravity theories. For example, if we consider a different coupling such as the Einstein-Maxwell-Dilaton theory [33, 34], the induced QNMs would exhibit different frequencies. Moreover, if we can imagine an astrophysical phenomenon where both effects are pronounced, it could present a valuable opportunity for multi-messenger astronomy[11].

Following [1], we briefly consider a sample source term within the Dirac delta potential approximation, detailed in section IV and section III of [1], and solve for the resulting second-order electromagnetic perturbation. If we assume a quadratic QNM source term, as in [35], we have

S=Φ(1)Ψ(1)(1+ζ|x|)2e(VEM2+VG2)(t|x|)(1+ζ|x|)2.S=\frac{\Phi^{(1)}\Psi^{(1)}}{(1+\zeta|x|)^{2}}\propto\frac{\,e^{-\left(\frac{V_{\text{EM}}}{2}+\frac{V_{G}}{2}\right)(t-|x|)}}{(1+\zeta|x|)^{2}}. (18)

Here we only focus on the QNM part. Typically, we should expect a correction to the amplitude of the electromagnetic linear modes and the excitation of new frequencies given by

Ω(2)=iVG+VEM2.\Omega^{(2)}=-i\frac{V_{G}+V_{EM}}{2}.

For x>0x>0 and x<0x<0, respectively, we have the second-order perturbation ΦQNM(2)\Phi^{(2)}_{QNM}:

ΦQNM(2)(t,x>0)=8ζ2VGeuVEM2{VEMeVEMζ[Ei(uVEM2+VEMζ)Ei(VEMζ)]+ζ}\displaystyle\Phi^{(2)}_{QNM}(t,x>0)=\frac{8}{\zeta^{2}V_{G}}e^{-\frac{uV_{\text{EM}}}{2}}\left\{V_{\text{EM}}e^{-\frac{V_{\text{EM}}}{\zeta}}\left[\operatorname{Ei}\left(\frac{uV_{\text{EM}}}{2}+\frac{V_{\text{EM}}}{\zeta}\right)-\operatorname{Ei}\left(\frac{V_{\text{EM}}}{\zeta}\right)\right]+\zeta\right\} (19)
+8ζ2VGeuVEM+VG2{(VEM+VG)eVEM+VGζ[Ei(VEM+VGζ)Ei((VEM+VG)(u2+1ζ))]ζ}\displaystyle\quad\quad\quad\quad+\frac{8}{\zeta^{2}V_{G}}e^{-u\frac{V_{\text{EM}}+V_{G}}{2}}\left\{(V_{\text{EM}}+V_{G})e^{-\frac{V_{\text{EM}}+V_{G}}{\zeta}}\left[\operatorname{Ei}\left(\frac{V_{\text{EM}}+V_{G}}{\zeta}\right)-\operatorname{Ei}\left((V_{\text{EM}}+V_{G})\left(\frac{u}{2}+\frac{1}{\zeta}\right)\right)\right]-\zeta\right\}
ΦQNM(2)(t,x<0)=8ζ2VGevVEM2{eVEMζVEM[Ei(uVEM2+VEMζ)Ei(VEMζ)]+ζ}\displaystyle\Phi^{(2)}_{QNM}(t,x<0)=\frac{8}{\zeta^{2}V_{G}}e^{-\frac{vV_{\text{EM}}}{2}}\left\{e^{-\frac{V_{\text{EM}}}{\zeta}}V_{\text{EM}}\left[\operatorname{Ei}\left(\frac{uV_{\text{EM}}}{2}+\frac{V_{\text{EM}}}{\zeta}\right)-\operatorname{Ei}\left(\frac{V_{\text{EM}}}{\zeta}\right)\right]+\zeta\right\} (20)
+8ζ2VGeuVG2vVEM2{(VEM+VG)eVEM+VGζ[Ei(VEM+VGζ)Ei((VEM+VG)(u2+1ζ))]ζ}\displaystyle\quad\quad\quad\quad+\frac{8}{\zeta^{2}V_{G}}e^{-\frac{uV_{G}}{2}-\frac{vV_{\text{EM}}}{2}}\left\{(V_{\text{EM}}+V_{G})e^{-\frac{V_{\text{EM}}+V_{G}}{\zeta}}\left[\operatorname{Ei}\left(\frac{V_{\text{EM}}+V_{G}}{\zeta}\right)-\operatorname{Ei}\left((V_{\text{EM}}+V_{G})\left(\frac{u}{2}+\frac{1}{\zeta}\right)\right)\right]-\zeta\right\}

III Perturbations due to point charges

In this section, we analyze the electromagnetic perturbation due to one or more point charges in Schwarzschild coordinates. Consider a particle moving along a worldline parameterized by an affine parameter τ\tau. The four-velocity vector UμU^{\mu} of the particle is given by

Uμ=(dtp(τ)dτ,dr¯p(τ)dτ,dθ¯p(τ)dτ,dϕ¯p(τ)dτ),U^{\mu}=\left(\frac{dt_{p}(\tau)}{d\tau},\frac{d\bar{r}_{p}(\tau)}{d\tau},\frac{d\bar{\theta}_{p}(\tau)}{d\tau},\frac{d\bar{\phi}_{p}(\tau)}{d\tau}\right), (21)

which can also be reparameterized in terms of the time coordinate tpt_{p} as:

Uμ=1dtpdτ(1,drp(tp)dtp,dθp(tp)dtp,dϕp(tp)dtp).U^{\mu}=\frac{1}{\frac{dt_{p}}{d\tau}}\left(1,\frac{dr_{p}(t_{p})}{dt_{p}},\frac{d\theta_{p}(t_{p})}{dt_{p}},\frac{d\phi_{p}(t_{p})}{dt_{p}}\right).

The four-current vector JμJ_{\mu} associated with the charge qq can then be written as:

Jμ\displaystyle J_{\mu} ={f(r),1f(r)drp(t)dt,r2dθp(t)dt,r2sin2θdϕp(t)dt}\displaystyle=\left\{f(r),-\frac{1}{f(r)}\frac{dr_{p}(t)}{dt},-r^{2}\frac{d\theta_{p}(t)}{dt},-r^{2}\sin^{2}\theta\frac{d\phi_{p}(t)}{dt}\right\} (22)
×qδ(rrp(t))r2lmYlm(θ,ϕ)Ylm(θp(t),ϕp(t)).\displaystyle\quad\times q\,\frac{\delta(r-r_{p}(t))}{r^{2}}\sum_{lm}Y_{lm}(\theta,\phi)Y^{*}_{lm}(\theta_{p}(t),\phi_{p}(t)).

This four-current vector can be further decomposed using vector spherical harmonics. The components of the current can be expressed as

Jμ\displaystyle J_{\mu} ={j(1)lmYlm,j(2)lmYlm,j(3)lmY˙lm+j(4)lmcscθYlm,\displaystyle=\left\{j_{(1)lm}Y_{lm},j_{(2)lm}Y_{lm},j_{(3)lm}\dot{Y}_{lm}+j_{(4)lm}\csc\theta\,Y^{\prime}_{lm},\right. (23)
j(3)lmYlmj(4)lmsinθY˙lm},\displaystyle\quad\left.j_{(3)lm}Y^{\prime}_{lm}-j_{(4)lm}\sin\theta\dot{Y}_{lm}\right\},

where the coefficients j(i)lmj_{(i)lm} are extracted by projecting over the sphere as outlined in D:

j(i)lm=𝑑Ω𝐘(i)μJμ.j_{(i)lm}=\int d\Omega\mathbf{Y}^{\mu*}_{(i)}J_{\mu}. (24)

Consequently,

j(1)lm\displaystyle j_{(1)lm} =qf(r)δ(rrp)r2Ylm(θp,ϕp),\displaystyle=qf(r)\frac{\delta(r-r_{p})}{r^{2}}\,Y^{*}_{lm}(\theta_{p},\phi_{p}), (25)
j(2)lm\displaystyle j_{(2)lm} =qf(r)drpdtδ(rrp)r2Ylm(θp,ϕp),\displaystyle=-\frac{q}{f(r)}\frac{dr_{p}}{dt}\frac{\delta(r-r_{p})}{r^{2}}Y^{*}_{lm}(\theta_{p},\phi_{p}), (26)
j(3)lm\displaystyle j_{(3)lm} =qλδ(rrp){imdϕpdtdθpdt𝒜lm[Ylm]},\displaystyle=\frac{q}{\lambda}\delta\left(r-r_{p}\right)\left\{im\frac{d\phi_{p}}{dt}-\frac{d\theta_{p}}{dt}\mathcal{A}_{lm}[Y_{lm}^{*}]\right\}, (27)
j(4)lm\displaystyle j_{(4)lm} =qλδ(rrp){imdθpdt𝒞lm[Ylm]+dϕpdt𝒟lm[Ylm]}.\displaystyle=\frac{q}{\lambda}\delta\left(r-r_{p}\right)\left\{im\frac{d\theta_{p}}{dt}\mathcal{C}_{lm}[Y_{lm}^{*}]+\frac{d\phi_{p}}{dt}\mathcal{D}_{lm}[Y_{lm}^{*}]\right\}. (28)

Here, 𝒜lm\mathcal{A}_{lm}, 𝒞lm\mathcal{C}_{lm}, and 𝒟lm\mathcal{D}_{lm} are operators that effectively act on the multipolar numbers of the spherical harmonics, inducing mulipolar mode-mixing defined as

𝒜lm[klm]\displaystyle\mathcal{A}_{lm}[k_{lm}] =lmklm𝑑ΩYlmYlmθ,\displaystyle=\sum_{lm}\,k_{lm}\int d\Omega Y_{lm}\frac{\partial Y_{l^{\prime}m^{\prime}}^{*}}{\partial\theta}\,, (29)
𝒞lm[klm]\displaystyle\mathcal{C}_{lm}[k_{lm}] =lmklm𝑑ΩcscθYlmYlm,\displaystyle=\sum_{lm}\,k_{lm}\int d\Omega\csc\theta\,Y_{lm}Y_{l^{\prime}m^{\prime}}^{*}\,,
𝒟lm[klm]\displaystyle\mathcal{D}_{lm}[k_{lm}] =lmklm𝑑ΩsinθYlmYlmθ.\displaystyle=\sum_{lm}\,k_{lm}\int d\Omega\sin\theta\,Y_{lm}\frac{\partial Y_{l^{\prime}m^{\prime}}}{\partial\theta}\,.

The effective source term for the Regge-Wheeler equation for multipoles l>0l>0 (even perturbations) Φe{}_{e}\Phi is given by:

𝒮elmJ=f(r)λ(r[r2j(1)lm]r2j(2)lmt).{}_{J}\mathcal{S}_{\text{e}\,lm}=\frac{f(r)}{\lambda}\left(\frac{\partial}{\partial r}\left[r^{2}j_{(1)lm}\right]-r^{2}\frac{\partial j_{(2)lm}}{\partial t}\right). (30)

For the point charge case, we obtain:

𝒮elmJ=\displaystyle{}_{J}\mathcal{S}_{\text{e}\,lm}= qλ{[r˙p(t)(θ˙p(t)Ylmθimϕ˙p(t)Ylm)\displaystyle\frac{q}{\lambda}\Bigg{\{}\Bigg{[}\,\dot{r}_{p}(t)\left(\dot{\theta}_{p}(t)\frac{\partial Y_{lm}^{*}}{\partial\theta}-im\,\dot{\phi}_{p}(t)Y_{lm}^{*}\right) (31)
+(f(r)f(r)dr+r¨p(t))Ylm]δ(rrp(t))\displaystyle+\left(f(r)\frac{f(r)}{dr}+\ddot{r}_{p}(t)\right)Y_{lm}^{*}\Bigg{]}\delta\left(r-r_{p}(t)\right)
+[f(r)2r˙p2(t)]δ(rrp(t))Ylm}.\displaystyle+\left[f(r)^{2}-\dot{r}_{p}^{2}(t)\right]\delta^{\prime}\left(r-r_{p}(t)\right)Y_{lm}^{*}\Bigg{\}}.

The odd perturbation Φo{}_{o}\Phi is described by

𝒮olmJ=f(r)j(4)lm.{}_{J}\mathcal{S}_{\text{o}\,lm}=-f(r)j_{(4)lm}. (32)

Subsequently, for the point particle case we have

𝒮olmJ=\displaystyle{}_{J}\mathcal{S}_{\text{o}\,lm}= qλf(r)δ(rrp(t)){imdθpdt𝒞lm[Ylm]\displaystyle-\frac{q}{\lambda}f(r)\delta\left(r-r_{p}(t)\right)\left\{im\frac{d\theta_{p}}{dt}\mathcal{C}_{lm}[Y_{lm}^{*}]\right. (33)
+dϕpdt𝒟lm[Ylm]}.\displaystyle\left.+\frac{d\phi_{p}}{dt}\mathcal{D}_{lm}[Y_{lm}^{*}]\right\}.

Finally, the general solution for Φlme,o{}_{e,o}\Phi_{lm} is given by

Φlme,o(r,t)=𝑑t2M𝑑rGlm(t,t,r,r)𝒮e,olmJ,{}_{e,o}\Phi_{lm}\left(r^{\prime},t^{\prime}\right)=\int_{-\infty}^{\infty}dt\int_{2M}^{\infty}dr\,G_{lm}\left(t,t^{\prime},r,r^{\prime}\right){}_{J}\mathcal{S}_{e,o\,lm}, (34)

where Glm(t,t,r,r)G_{lm}(t,t^{\prime},r,r^{\prime}) represents the Green’s function for the Regge-Wheeler master equation with spin s=1s=1, see Section II in [1] for further details. For completeness, we emphasize that for any signal observed at a specific event (t,x)(t,x), the Green’s functions define a limited domain of support in the (t,x)(t^{\prime},x^{\prime}) plane [36, 37]. There are 2D sub-regions in the (t,x)(t^{\prime},x^{\prime}) space that contribute to the event at (t,x)(t,x) through convolution of the Green’s function with the source term 𝒮e,olm\mathcal{S}_{e,o\,lm}. In the case of a point particle, this contribution is restricted to a 1D curve in (t,x)(t^{\prime},x^{\prime}). Consequently, the problem reduces to a path integral along the worldline of the particle, which can be reparameterized over a segment of the path with respect to any variable that is one-to-one with the affine parameter on that segment.

For example, if we stick to tpt_{p} as affine parameter, then for odd perturbation Φlmo{}_{o}\Phi_{lm} the solution becomes

Φlmo(r,t)=\displaystyle{}_{o}\Phi_{lm}(r^{\prime},t^{\prime})= qλdtpf(rp(tp)){imdθpdtp𝒞lm[Ylm]\displaystyle-\frac{q}{\lambda}\int_{-\infty}^{\infty}dt_{p}\,f\left(r_{p}(t_{p})\right)\left\{im\frac{d\theta_{p}}{dt_{p}}\mathcal{C}_{lm}[Y_{lm}^{*}]\right. (35)
+dϕpdtp𝒟lm[Ylm]}Glm(r,t,rp(tp),tp).\displaystyle\left.+\frac{d\phi_{p}}{dt_{p}}\mathcal{D}_{lm}[Y_{lm}^{*}]\right\}\,G_{lm}\left(r^{\prime},t^{\prime},r_{p}(t_{p}),t_{p}\right).

For the even perturbation Φlme{}_{e}\Phi_{lm}

e Φlm(t,r)=qλdtp{[r¨p(tp)f(rp(tp))df(rp)drp\displaystyle\Phi_{lm}\left(t^{\prime},r^{\prime}\right)=\frac{q}{\lambda}\int_{-\infty}^{\infty}dt_{p}\Bigg{\{}\left[\ddot{r}_{p}(t_{p})-f\left(r_{p}(t_{p})\right)\frac{df\left(r_{p}\right)}{dr_{p}}\right. (36)
imdrpdtpdϕpdtp+mcotθpdrpdtpdθpdtp]Ylm\displaystyle\left.-im\,\frac{dr_{p}}{dt_{p}}\frac{d\phi_{p}}{dt_{p}}+m\cot\theta_{p}\,\frac{dr_{p}}{dt_{p}}\frac{d\theta_{p}}{dt_{p}}\right]Y_{lm}^{*}
+[(lm)(l+m+1)eiϕ^drpdtpdθpdtp]Yl,m+1}\displaystyle+\left[\sqrt{(l-m)(l+m+1)}e^{i\hat{\phi}}\frac{dr_{p}}{dt_{p}}\frac{d\theta_{p}}{dt_{p}}\right]Y_{l,m+1}^{*}\Bigg{\}}
×Glm(r,t,rp(tp),tp)+qλ𝑑tpGlmdr|r=rp(tp)\displaystyle\times G_{lm}\left(r^{\prime},t^{\prime},r_{p}(t_{p}),t_{p}\right)+\frac{q}{\lambda}\int^{\infty}_{-\infty}dt_{p}\left.\frac{G_{lm}}{dr}\right|_{r=r_{p}(t_{p})}
×[r˙p(tp)2f2(rp(tp))]Ylm.\displaystyle\times\left[\dot{r}_{p}(t_{p})^{2}-f^{2}\left(r_{p}(t_{p})\right)\right]Y_{lm}^{*}.

However, as we will see in the upcoming subsections, a well-chosen parameter will depend on the worldline being examined, and possibly on the (t,x)(t,x) coordinates in which we are aiming to examine the perturbations. A thoughtful choice of this parameter can significantly simplify the complexity of the computation.

III.1 Ideal dipole

The linearity of Maxwell’s equations extends naturally to the electromagnetic Regge-Wheeler equation. If the total current JμJ_{\mu} is the sum of the currents from two particles, i.e.,

Jμ=Jμparticle(1)+Jμparticle(2),J_{\mu}=J_{\mu}^{\text{\scriptsize particle(1)}}+J_{\mu}^{\text{\scriptsize particle(2)}}, (37)

then the resulting scalar field Φ\Phi is the superposition of the fields from each particle:

Φ=Φparticle(1)+Φparticle(2).\Phi=\Phi^{\text{\scriptsize particle(1)}}+\Phi^{\text{\scriptsize particle(2)}}. (38)

Here, Φparticle(1)\Phi^{\text{\scriptsize particle(1)}} and Φparticle(2)\Phi^{\text{\scriptsize particle(2)}} are the solutions to the Regge-Wheeler equation for each current. For a simple dipole structure, consider a particle with charge qq on worldline γ1\gamma_{1} parameterized by rp1r_{p1}, and another with charge q-q on worldline γ2\gamma_{2} parameterized by rp2r_{p2}, where rp2rp1η+𝒪(η2)r_{p2}-r_{p1}\approx\eta+\mathcal{O}(\eta^{2})222In case of radial free fall of a radial dipole 𝒪(η2)\mathcal{O}(\eta^{2})=0 and the expansion truncates at first order correction in η\eta.. Thus, the total field for the dipole could simply be written as a variation of a the perturbation due to a positive charge as

Φde,o=δΦ+e,o.{}_{e,o}\Phi^{d}=\delta\,{}_{e,o}\Phi^{+}. (39)

The complexity of solving this problem can be divided into three parts:

  1. 1.

    Solving for the worldines of charges qq and q-q considering the electromagnetic force between the particles and self-interaction.

  2. 2.

    Solving for the Green’s function GlmG_{lm}.

  3. 3.

    Evaluating the one-dimensional integral to obtain Φe,o{}_{e,o}\Phi.

We will now consider some simple cases.

III.2 Examples

III.2.1 Point charge in radial free fall

Consider a charge qq freely falling radially from rest with E=1E=1 and with angular coordinate (θ0,ϕ0)(\theta_{0},\phi_{0}), while drpdtp=β(rp)f(rp)\frac{dr_{p}}{dt_{p}}=-\beta(r_{p})f(r_{p}) and j(3)lm=j(4)lm=0j_{(3)lm}=j_{(4)lm}=0 vanishes. We have

Φlmo(t,r)=0.\displaystyle{}_{o}\Phi_{lm}\left(t^{\prime},r^{\prime}\right)=0. (40)

After re-paramterization of the integral in terms of rpr_{p}, as outlined in Appendix C, the even part is

Φlme(t,r)=qYlmλ[\displaystyle{}_{e}\Phi_{lm}\left(t^{\prime},r^{\prime}\right)=\frac{-qY_{lm}^{*}}{\lambda}\Bigg{[} drp{(3f(rp)2β(rp)df(rp)drp)Glm\displaystyle\int dr_{p}\Bigg{\{}\left(-\frac{3f(r_{p})}{2\beta(r_{p})}\frac{df(r_{p})}{dr_{p}}\right)G_{lm} (41)
+dGlmdr|r=rp(f(rp)2β(rp))}].\displaystyle+\left.\frac{dG_{lm}}{dr}\right|_{r=r_{p}}\left(-\frac{f(r_{p})^{2}}{\beta(r_{p})}\right)\Bigg{\}}\Bigg{]}.

III.2.2 Point charge in a circular orbit

Consider a charge qq in a circular orbit at a constant radius rpr_{p}, with angular coordinates (θ0,ϕp(tp))(\theta_{0},\phi_{p}(t_{p})). The angular position ϕp(tp)\phi_{p}(t_{p}) is related to time tpt_{p} by tp=ϕpΩ¯t_{p}=\frac{\phi_{p}}{\bar{\Omega}}, where Ω¯\bar{\Omega} is the constant angular velocity of the particle, and j(2)lm=0j_{(2)lm}=0. The odd-parity perturbation is given by

Φlmo(t,r)=qλimf(r0)\displaystyle{}_{o}\Phi_{lm}\left(t^{\prime},r^{\prime}\right)=\frac{q}{\lambda}imf(r_{0}) ϕpiϕpf𝑑ϕp𝒟lm[Ylm]θ=θ0,ϕ=ϕp\displaystyle\int_{\phi_{pi}}^{\phi_{pf}}d\phi_{p}\,\mathcal{D}_{lm}\left[Y_{lm}^{*}\right]_{\theta=\theta_{0},\phi=\phi_{p}} (42)
×Glm(r,t,r0,ϕpΩ).\displaystyle\times G_{lm}\left(r^{\prime},t^{\prime},r_{0},\frac{\phi_{p}}{\Omega}\right).

The even-parity perturbation is given by

Φlme(t,r)=qλΩϕpiϕpf𝑑ϕpYlm(θ0,ϕp)\displaystyle{}_{e}\Phi_{lm}\left(t^{\prime},r^{\prime}\right)=\frac{-q}{\lambda\Omega}\int_{\phi_{pi}}^{\phi_{pf}}d\phi_{p}\,Y^{*}_{lm}(\theta_{0},\phi_{p}) (43)
×[ff(r)r|r=r0Glm(t,r,ϕpΩ,r0)+f2Glmr|r=r0]\displaystyle\quad\times{\left[f\left.\frac{\partial f(r)}{\partial r}\right|_{r=r_{0}}G_{lm}(t^{\prime},r^{\prime},\frac{\phi_{p}}{\Omega},r_{0})+f^{2}\left.\frac{\partial G_{lm}}{\partial r}\right|_{r=r_{0}}\right]}

III.2.3 Quasi-circular orbit

We can imagine a worldline where the particle’s motion is confined to a plane at θ=π2\theta=\frac{\pi}{2} while undergoing angular motion in the ϕ^\hat{\phi} direction, similar to that of a circular orbit. Simultaneously, the particle experiences radial motion with velocity vv, where vrΩ^v\ll r\hat{\Omega}, starting from rinnerr_{\text{inner}} and moving to routerr_{\text{outer}}. We can utilize integrals (36,35) with drpdtp\frac{dr_{p}}{dt_{p}} and dϕpdtp\frac{d\phi_{p}}{dt_{p}} as constants, while dθpdtp=0\frac{d\theta_{p}}{dt_{p}}=0. This approach approximates the quasi-circular worldline of the point particle.

IV Dirac delta function potential

Away from the light ring, the potential diminishes, and solutions become asymptotically outgoing (or ingoing) near the flat region (or horizon), which is approximately true for the Regge-Wheeler potential at large distances. This holds analytically for superlocalized potentials like the Dirac delta function VEMδ(x)V_{EM}\delta(x)[35]. The Green’s function333See [38, 35, 3] for a detailed discussion of the flat Green’s function component GFG_{F}, the QNM component GQNMG_{QNM}, and the branch cut component GBG_{B}. is expressed as:

G(tt,r,r)=GF(tt,r,r)+GQ(tt,r,r),\begin{gathered}G(t-t^{\prime},r,r^{\prime})=G_{F}(t-t^{\prime},r,r^{\prime})+G_{Q}(t-t^{\prime},r,r^{\prime}),\end{gathered} (44)

The Green’s function is further written as

G(T,r,r)=12[Θ(A)Θ(B)]12eVEM2BΘ(B),G(T,r,r^{\prime})=-\frac{1}{2}\left[\Theta(A)-\Theta(B)\right]-\frac{1}{2}e^{-\frac{V_{EM}}{2}B}\Theta(B), (45)

where Θ\Theta is the Heaviside step function. The derivative of the Green’s function with respect to rr^{\prime} is:

f(r)\displaystyle f(r^{\prime}) G(tt,r,r′′)r′′|r′′=r=12[sδ(A)s~δ(B)]\displaystyle\frac{\partial\,G(t-t^{\prime},r,r^{\prime\prime})}{\partial r^{\prime\prime}}\bigg{|}_{r^{\prime\prime}=r^{\prime}}=\frac{1}{2}\left[s\delta(A)-\tilde{s}\delta(B)\right] (46)
s~2eVEM2(B)[VEM2Θ(B)δ(B)],\displaystyle\quad-\frac{\tilde{s}}{2}e^{-\frac{V_{EM}}{2}(B)}\left[\frac{V_{EM}}{2}\Theta(B)-\delta(B)\right],
444Taking derivative of with Dirac delta function distribution should be done under the integration sign, that is why we are not canceling the two contributions of δ(B)\delta(B), in case higher derivatives are needed.

Here, s(xx)s(x-x^{\prime}) and s~(xxEM)\tilde{s}(x-x_{EM}) are sign functions, taking ++ for positive arguments and - for negative arguments. The terms AA and BB encode the causal structure, defined as:

A\displaystyle A =tt|xx|,\displaystyle=t-t^{\prime}-|x-x^{\prime}|, (47)
B\displaystyle B =tt|x||x|,\displaystyle=t-t^{\prime}-|x|-|x^{\prime}|,

IV.1 Support regions

The causal structure becomes more evident and sharper with simpler potentials, such as the Dirac delta potential. In this case, the QNMs component of the Green’s function is supported by the past light cone of the point (0,tx)(0,t-x), which is defined by the region BB in the (t,x)(t^{\prime},x^{\prime}) plane.

Focusing on point sources, a particle moving along a trajectory γ\gamma contributes to the QNM signal observed at (x,t)(x,t) primarily due to the particle’s interaction with the potential along its worldline. Almost all of this contribution arises from the particle’s history in proximity of the potential. Specifically, if the particle starts from an asymptotically far region, rr\rightarrow\infty, its contribution begins at tt^{\prime}\rightarrow-\infty and ends when its trajectory exits the past light cone of the point (0,tx)(0,t-x) in the (t,x)(t^{\prime},x^{\prime}) plane.

In many cases, as seen in the literature [39, 35], the source’s contribution is considered to start at a finite time, particularly when gravitational QNMs source the quadratic QNMs, as in binary black hole merger scenarios. While starting the particle’s contribution at tt^{\prime}\rightarrow-\infty helps get rid of contribution from electromagnetic field I.C., it can be helpful to consider the particle (the source) starting its contribution at a finite time at some large radius r0r_{0}, to simplify the analysis. By setting the beginning of time to t=0t=0, the source’s contribution is defined by its intersection with the region BB, bounded below by t=0t^{\prime}=0 in the (t,x)(t^{\prime},x^{\prime}) space. For further clarification, visual references are useful; see Figures 2 (a–f).

In the Figures 2, we imagine a point charge qq described by the solid green curve contributing to the signal observed at x=3Mx=3M. Initially, in (a), there is no QNM contribution at t=10Mt=10M, but as time progresses as shown in (b) and (c), the support region BB expands, capturing more of the particle’s worldline as it enters the region. In contrast, when the observation point xx gets further away from the potential (as shown in (f)) the support region shrinks down. In that regime, in the limit of xx\rightarrow\infty, the contribution only near the potential will excite QNM signal. If we assume that the particle started from rr\rightarrow\infty while tt\rightarrow-\infty, the worldline would always remain inside the region BB of any (t,x)(t,x), and these past light cones BB would extend to infinity.

For GQG_{Q}, the redness of the contours, increases on a light-like mesh, with inner past light cones appearing whiter than the outer ones. The outermost part represents the edge of the signal, while the innermost part corresponds to the signal’s end.

The integration limits are (t,x)(t,x)-dependent. As we have seen before, if we choose to parameterize the path integral, defined by the overlap of the green curve with the BB region (colored in red), using the parameter rpr_{p} for certain geodesics555i.e. radial free fall the integration limits are rpBmax(t,x)r_{pBmax}(t,x) and rpBmin(t,x)r_{pBmin}(t,x). For a particle starting at infinity, we generally have rpBmin(t,x)r_{pBmin}(t,x)\rightarrow\infty. However, solving for rpBmax(t,x)r_{pBmax}(t,x) requires solving for B=0B=0 in terms of rpr_{p}, given by

t|x|=t(rp)+|x(rp)|,t-|x|=t^{\prime}(r_{p})+|x^{\prime}(r_{p})|,

ensuring the integration is properly constrained and accounts for the causal structure encoded in BB.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Refer to caption
(e)
Refer to caption
(f)
Figure 2: The red region represents GQG_{Q}, the QNM support region, with redness indicating the weight of GQG_{Q} (where bright red signifies lower weight). For x>0x>0, the brown region denotes GFG_{F}, the flat support region, while for x<0x<0, GFG_{F} is shown in cyan. Figures (a), (b), and (c) illustrate the expansion of the support regions over time at a fixed observation distance of x=3Mx=3M, shown at t=10Mt=10M, t=20Mt=20M, and t=30Mt=30M, respectively. Comparing (b), (e), and (f) highlights the support regions’ shrinking effect as xx increases from 3M3M to 9M9M to 15M15M for the same moment in time 20M20M. As the observation distance xx increases, the QNM support region becomes more localized around the potential, leading to an asymptotic observer perceiving an almost constant amplitude. Though the amplitude generally depends on both tt and xx, this sensitivity diminishes with increased distance as the support region becomes concentrated near the potential. Comparison between (d) and (f) highlights the flat regions GFG_{F} at x=15Mx=15M (in brown) and x=15Mx=-15M (in cyan). The solid green and dashed green lines represent point charges +q+q and q-q, separated by η=0.6M\eta=0.6M, with their initial position at r0=12.7Mr_{0}=12.7M and r0+η=13.3Mr_{0}+\eta=13.3M. The asymmetry in the worldlines of these charges illustrates that, for both flat and QNM parts, for x<0x<0 (inside the potential at x=0x=0), contributions are delayed compared to events at the same distance x>0x>0 (outside the potential).

To study the flat contribution of the Green’s function, the situation is analogous except that support region is ABA-B, where AA represents another past light cone centered at (x,t)(x,t). The reader can observe that the flat part’s density plot, as in figures 2, is constant and colored in brown for x>0x>0 (or cyan for x<0x<0), as the value of GFG_{F} inside ABA-B remains a constant =1/2=-1/2.

Another contribution we should account for is what, mathematically speaking, can be regarded as the evaluation of the derivative of a Dirac delta distribution under the integration sign. Physically, we can interpret this derivative as an additional contribution arising from the edges of the QNM region BB and the flat region ABA-B, where the particle is about to penetrate those regions. This contribution should manifest at the edge of the signal. Hence, we will refer only on the integration with GQG_{Q} as the QNM part.

Moreover, the problem is not fully symmetric around the potential peak x=0x=0. Although the Green’s function itself is symmetric, our particular source term, the worldline of the particle, freezes in proximity to the horizon xx\rightarrow-\infty. Therefore, only at late times, for equal Δx\Delta x around the potential near the horizon or near the asymptotic flat region, will a QNM be observed. The same holds true for flat contributions, see (d) and (f) for comparison.

V First-order electromagnetic perturbation due to point charges

We are now in a good position to proceed with solving for the electromagnetic perturbation for radial free fall case. Generically, the path integrals in III.2 may present challenges; therefore, the choice between a semi-analytical approximation or a numerical solution depends on which aspects of the solution are of primary interest.

V.1 Point charge in radial free fall

For simplicity, we assume homogeneous initial conditions on Φlm(1)e(r,t){}_{e}\Phi^{(1)}_{lm}(r,t):

Φlm(1)e(r,t)\displaystyle{}_{e}\Phi^{(1)}_{lm}(r,t) =qλYlm(θ0,ϕ0)Φ^(1)(t,r),\displaystyle=\frac{-q}{\lambda}Y_{lm}^{*}(\theta_{0},\phi_{0})\hat{\Phi}^{(1)}(t,r), (48)
Φ^(1)(t,r)\displaystyle\hat{\Phi}^{(1)}(t,r) =I1(t,r)+I2(t,r),\displaystyle=I_{1}(t,r)+I_{2}(t,r),

it’s important to note that Φ^(1)(t,r)\hat{\Phi}^{(1)}(t,r) in this particular case is independent of the multipole numbers (l,m)(l,m). The integrals I1(t,r)I_{1}(t,r) and I2(t,r)I_{2}(t,r) are defined as follows666where Θ(t|x|)\Theta(t-|x|) could be factored out just for clarity, although it should still be embedded in the BB integrals regardless:

I1\displaystyle I_{1} =32𝑑tGf(rp(t))2df(rp(t))drp(t)\displaystyle=\frac{3}{2}\int_{-\infty}^{\infty}dt^{\prime}\,G\,f(r_{p}(t^{\prime}))^{2}\frac{df(r_{p}(t^{\prime}))}{dr_{p}(t^{\prime})}
34[IF1AΘIF1B+ΘIQ1BeVEM2(t|x|)],\displaystyle\equiv\frac{3}{4}\bigg{[}I_{F1A}-\Theta I_{F1B}+\Theta I_{Q1B}e^{-\frac{V_{EM}}{2}(t-|x|)}\bigg{]}, (49)

and

I2\displaystyle I_{2} =𝑑tGr|r=rp(t)f(rp(t))3\displaystyle=\int_{-\infty}^{\infty}dt^{\prime}\,\frac{\partial G}{\partial r^{\prime}}\bigg{|}_{r^{\prime}=r_{p}(t^{\prime})}f(r_{p}(t^{\prime}))^{3}
12[sIF2AΘs~VEM2eVEM2(t|x|)IQ2B].\displaystyle\equiv-\frac{1}{2}\bigg{[}sI_{F2A}-\Theta\tilde{s}\frac{V_{EM}}{2}e^{-\frac{V_{EM}}{2}(t-|x|)}I_{Q2B}\bigg{]}. (50)

The “amplitude” A(t,x)A(t,x) of the QNM part can be expressed as

A(t,x)=34IQ1B+s~4VEMIQ2B,A(t,x)=\frac{3}{4}I_{Q1B}+\frac{\tilde{s}}{4}V_{EM}I_{Q2B},

However, evaluating IQ1BI_{Q1B} and IQ2BI_{Q2B} may not be as straightforward as the other integrals, as the reader can see in Appendix B. While these integrals can be computed numerically, if we solve for rpBmax(t,x)r_{pBmax}(t,x) numerically. For simplicity, we will proceed with I.C. on the trajectory that t=0t=0 at rpBmin=r0r_{pBmin}=r_{0} 777if we are interested in the case starting far from the black hole, i.e. r0r_{0}\rightarrow\infty, this will require that tt runs <t<-\infty<t<\infty., and also that tt is positive definite. Since the integration limits are often easier to handle in double null coordinates uu^{\prime} and vv^{\prime}, we can proceed with the following transformation of variables in the right half on the (t,x)(t^{\prime},x^{\prime}) plane as

v\displaystyle v v(rp)=t(rp)+x(rp),\displaystyle\mapsto v(r_{p})=t(r_{p})+x(r_{p}),
dv\displaystyle-dv f(1+β)=fβdrp,\displaystyle\,f(1+\beta)=\frac{f}{\beta}dr_{p},

while for the left half we have

u\displaystyle u u(rp)=t(rp)x(rp),\displaystyle\mapsto u(r_{p})=t(r_{p})-x(r_{p}),
du\displaystyle-du f(1β)=fβdrp.\displaystyle\,f(1-\beta)=\frac{f}{\beta}dr_{p}.

The integration limits are a bit tricky as we encounter three different cases:

  1. 1.

    If we are interested in the case where the particle starts from a finite position x(r0)>t|x|x^{\prime}(r_{0})>t-|x|, then particle never enters the past light cone of (0,t|x|)(0,t-|x|). In this case, it would have zero QNM contribution for (t,x)(t,x).

  2. 2.

    The particle exits the past light cone before reaching the potential peak at x=0x=0. There is no contribution from left half of the light cone.

    rpBmin(t,r)=r0v=x(r0)rpBmax(t,r)v=t|x|.\begin{gathered}r_{pBmin}(t,r)=r_{0}\mapsto v=x(r_{0})\\ r_{pBmax}(t,r)\mapsto v=t-|x|.\end{gathered} (51)
    A2(t,x)=t|x|x(r0)𝑑v(β+1)f(34dfdrp+1~4VEM)evVEM2,A_{2}(t,x)=\int_{t-|x|}^{x(r_{0})}dv^{\prime}\,(\beta+1)f\left(\frac{3}{4}\frac{df}{dr_{p}}+\frac{\tilde{1}}{4}V_{EM}\right)e^{v^{\prime}\frac{V_{EM}}{2}}, (52)
  3. 3.

    The particle exits after passing x=0x=0 the past light cone. Then right side contribution in vv^{\prime}:

    rpBmin(t,r)=r0v=x(r0)v=t′′,\begin{gathered}r_{pBmin}(t,r)=r_{0}\mapsto v=x(r_{0})\\ \quad\quad v=t^{\prime\prime},\end{gathered} (53)

    where t′′t^{\prime\prime} represents the time at which the particle passes the peak of the potential at x=0x=0. The left side contribution in uu^{\prime} is

    u=t′′.rpBmax(t,r)v=t|x|.\begin{gathered}u=t^{\prime\prime}.\\ r_{pBmax(t,r)}\mapsto v=t-|x|.\end{gathered} (54)

    Hence, the integral should be divided into two halves before and after x=0x=0

    A3(t,x)\displaystyle A_{3}(t,x) =t|x|t′′𝑑u(β1)f(34dfdrp+s~4VEM)euVEM2\displaystyle=\int_{t-|x|}^{t^{\prime\prime}}du^{\prime}\,(\beta-1)f\left(\frac{3}{4}\frac{df}{dr_{p}}+\frac{\tilde{s}}{4}V_{EM}\right)e^{u^{\prime}\frac{V_{EM}}{2}} (55)
    +t′′x(r0)𝑑v(β+1)f(34dfdrp+s~4VEM)evVEM2.\displaystyle+\int_{t^{\prime\prime}}^{x(r_{0})}dv^{\prime}\,(\beta+1)f\left(\frac{3}{4}\frac{df}{dr_{p}}+\frac{\tilde{s}}{4}V_{EM}\right)e^{v^{\prime}\frac{V_{EM}}{2}}.

While the integration limits are now straightforward, this has resulted in a complex integrand, as was expected since the worldline is timelike. In other words, because it’s not easy to invert the function u(rp)u^{\prime}(r_{p}) (or v(rp)v^{\prime}(r_{p})) analytically. However, for some point (t,x)(t,x) such that t|x|t\approx|x|, we can expand around u=0u=0888When x>0x>0, we can use uu, for x<0x<0, we use vv. Physically we can find the initial behavior of the QNM at some position xx, however as tt increases, higher terms in the expansions need to be considered. A Taylor expansion of the segment of the function of rp(u)r_{p}(u) around u=0u=0 up to the order nn can be integrated as follows

Pn(u)euVEM2=euVEM2i=0ncik=0i(1)ki!(ik)!uik(VEM2)k,\int P_{n}(u)e^{u\frac{V_{EM}}{2}}=e^{u\frac{V_{EM}}{2}}\sum_{i=0}^{n}c_{i}\sum_{k=0}^{i}(-1)^{k}\frac{i!}{(i-k)!}\frac{u^{i-k}}{\left(\frac{V_{EM}}{2}\right)^{k}},

where cic_{i} are the coefficient of the expansion

(β+1)f(34dfdrp+s~4VEM)(\beta+1)f\left(\frac{3}{4}\frac{df}{dr_{p}}+\frac{\tilde{s}}{4}V_{EM}\right)

around u=0u=0 that could easily be obtained through applying the chain rule.

Back to case (2), evaluating the integrand above at the upper limit x(r0)x(r_{0}) will contribute to the QNM part with amplitude that only depends on the I.C. of the worldline. However the lower limit t|x|t-|x| will result in a flat contribution which is (t,x)(t,x) dependent.

eVEM2(t|x|)A2(t,x)=eVEM2(t|x|)𝐀2(r0)i=0nci(k=0i(1)ki!(ik)!(t|x|)ik(VEM2)k),\displaystyle e^{-\frac{V_{EM}}{2}(t-|x|)}A_{2}(t,x)=e^{-\frac{V_{EM}}{2}(t-|x|)}\mathbf{A}_{2}(r_{0})-\sum_{i=0}^{n}c_{i}\left(\sum_{k=0}^{i}(-1)^{k}\frac{i!}{(i-k)!}\frac{(t-|x|)^{i-k}}{\left(\frac{V_{EM}}{2}\right)^{k}}\right), (56)
𝐀2(r0)=ex(r0)VEM2i=0nci(k=0i(1)ki!(ik)!x(r0)ik(VEM2)k).\displaystyle\mathbf{A}_{2}(r_{0})=e^{x(r_{0})\frac{V_{EM}}{2}}\sum_{i=0}^{n}c_{i}\left(\sum_{k=0}^{i}(-1)^{k}\frac{i!}{(i-k)!}\frac{x(r_{0})^{i-k}}{\left(\frac{V_{EM}}{2}\right)^{k}}\right).

The case (3) should be very similar to the case (2), and we will end up with QNM part with amplitude only dependent on the initial conditions for the particle and a flat part polynomial in t|x|t-|x|. However, the coefficients of expansion will be different.

eVEM2(t|x|)A3(t,x)=eVEM2(t|x|)𝐀3(r0)i=0nfi(k=0i(1)ki!(ik)!(t|x|)ik(VEM2)k),\displaystyle e^{-\frac{V_{EM}}{2}(t-|x|)}A_{3}(t,x)=e^{-\frac{V_{EM}}{2}(t-|x|)}\mathbf{A}_{3}(r_{0})-\sum_{i=0}^{n}f_{i}\left(\sum_{k=0}^{i}(-1)^{k}\frac{i!}{(i-k)!}\frac{(t-|x|)^{i-k}}{\left(\frac{V_{EM}}{2}\right)^{k}}\right), (57)
𝐀3(r0)=ex(r0)VEM2i=0ndi(k=0i(1)ki!(ik)!x(r0)ik(VEM2)k)et′′VEM2i=0ndi(k=0i(1)ki!(ik)!t′′ik(VEM2)k)\displaystyle\mathbf{A}_{3}(r_{0})=e^{x(r_{0})\frac{V_{EM}}{2}}\sum_{i=0}^{n}d_{i}\left(\sum_{k=0}^{i}(-1)^{k}\frac{i!}{(i-k)!}\frac{x(r_{0})^{i-k}}{\left(\frac{V_{EM}}{2}\right)^{k}}\right)-e^{t^{\prime\prime}\frac{V_{EM}}{2}}\sum_{i=0}^{n}d_{i}\left(\sum_{k=0}^{i}(-1)^{k}\frac{i!}{(i-k)!}\frac{t^{\prime\prime\,i-k}}{\left(\frac{V_{EM}}{2}\right)^{k}}\right)
+et′′VEM2i=0nfi(k=0i(1)ki!(ik)!t′′ik(VEM2)k).\displaystyle+e^{t^{\prime\prime}\frac{V_{EM}}{2}}\sum_{i=0}^{n}f_{i}\left(\sum_{k=0}^{i}(-1)^{k}\frac{i!}{(i-k)!}\frac{t^{\prime\prime\,i-k}}{\left(\frac{V_{EM}}{2}\right)^{k}}\right).

Finally, the Maxwell tensor FF describing this perturbation can be written as

F=Fmonopole+Frade1+Finte2,F=F_{\text{monopole}}+{}_{e1}F_{\text{rad}}+{}_{e2}F_{\text{int}}, (58)

where the monopole contribution as well as the radiative and “internal” contribution are expressed in terms of Φ(1)\Phi^{(1)} and the current. The monopole contribution is simply Fmonopoletr=qΘ(rrp(t))r2F_{\text{monopole}\,tr}=q\frac{\Theta(r-r_{p}(t))}{r^{2}}. The radiative contribution is

Frade1=q(0[δ2(θ,θ0,ϕ,ϕ0)14π]Φ^(1)r20GS2θfΦ^(1)rGS2θ1fΦ^(1)t00GS2ϕfΦ^(1)rGS2ϕ1fΦ^(1)t00).\begin{aligned} {}_{e1}F_{\text{rad}}=q\begin{pmatrix}0&**&**&**\\ \left[\delta^{2}(\theta,\theta_{0},\phi,\phi_{0})-\frac{1}{4\pi}\right]\frac{\hat{\Phi}^{(1)}}{r^{2}}&0&**&**\\ -\frac{\partial G_{S^{2}}}{\partial\theta}f\frac{\partial\hat{\Phi}^{(1)}}{\partial r}&-\frac{\partial G_{S^{2}}}{\partial\theta}\frac{1}{f}\frac{\partial\hat{\Phi}^{(1)}}{\partial t}&0&0\\ -\frac{\partial G_{S^{2}}}{\partial\phi}f\frac{\partial\hat{\Phi}^{(1)}}{\partial r}&-\frac{\partial G_{S^{2}}}{\partial\phi}\frac{1}{f}\frac{\partial\hat{\Phi}^{(1)}}{\partial t}&0&0\end{pmatrix}\end{aligned}. (59)

While internal contribution is given by

Finte2=q(0000GS2θfδ(rrp(t))GS2θβδ(rrp(t))00GS2ϕfδ(rrp(t))GS2ϕβδ(rrp(t))00).\displaystyle{}_{e2}F_{\text{int}}=q\begin{pmatrix}0&0&**&**\\ 0&0&**&**\\ -\frac{\partial G_{S^{2}}}{\partial\theta}f\delta(r-r_{p}(t))&-\frac{\partial G_{S^{2}}}{\partial\theta}\beta\,\delta(r-r_{p}(t))&0&0\\ -\frac{\partial G_{S^{2}}}{\partial\phi}f\delta(r-r_{p}(t))&-\frac{\partial G_{S^{2}}}{\partial\phi}\beta\,\delta(r-r_{p}(t))&0&0\end{pmatrix}. (60)

V.2 Radial ideal dipole in radial free fall

For simplicity, we will focus on the scenario of radial free fall, and the radial dipole. Additionally, we will disregard the effects of electromagnetic interactions and self-interactions between the two particles along their worldlines. Instead, we consider two charges, qq and q-q, with a separation η\eta along their radial worldlines into the black hole, starting from rest at large distance positions r0r_{0} and r0+ηr_{0}+\eta respectively with E=1E=1.

We can derive expressions for an ideal radial dipole in the limit as η0\eta\rightarrow 0 as a variation of this expression with respect to r0r_{0} or equivalently with respect to t′′t^{\prime\prime}, denoted as δA(t,x)\delta A(t,x). Focusing on the QNM contribution, we will proceed with similar cases as we had before. In the dipole case (2) only segments of wordline xp>0x_{p}>0 contribute to the trajectory

1ηδA2δr0=[(β+1)f(34dfdrp+s~4VEM)evVEM2]r=r0+t|x|x(r0)𝑑vddrp[(β+1)f(34dfdrp+s~4VEM)]evVEM2.\displaystyle\frac{1}{\eta}\frac{\delta A_{2}}{\delta r_{0}}=\left[(\beta+1)f\left(\frac{3}{4}\frac{df}{dr_{p}}+\frac{\tilde{s}}{4}V_{EM}\right)e^{v^{\prime}\frac{V_{EM}}{2}}\right]_{r=r_{0}}+\int_{t-|x|}^{x(r_{0})}dv^{\prime}\,\frac{d}{dr_{p}}\left[(\beta+1)f\left(\frac{3}{4}\frac{df}{dr_{p}}+\frac{\tilde{s}}{4}V_{EM}\right)\right]e^{v^{\prime}\frac{V_{EM}}{2}}. (61)
1ηδA3δr0\displaystyle\frac{1}{\eta}\frac{\delta A_{3}}{\delta r_{0}} =r03/22(r02)[(β1)f(34dfdrp+s~4VEM)euVEM2(β+1)f(34dfdrp+s~4VEM)evVEM2]r=r0\displaystyle=\frac{r_{0}^{3/2}}{2(r_{0}-2)}\left[(\beta-1)f\left(\frac{3}{4}\frac{df}{dr_{p}}+\frac{\tilde{s}}{4}V_{EM}\right)e^{u^{\prime}\frac{V_{EM}}{2}}-(\beta+1)f\left(\frac{3}{4}\frac{df}{dr_{p}}+\frac{\tilde{s}}{4}V_{EM}\right)e^{v^{\prime}\frac{V_{EM}}{2}}\right]_{r=r_{0}} (62)
f(r0)[(β+1)f(34dfdrp+s~4VEM)evVEM2]r=r0+t|x|t′′𝑑uddrp[(β1)f(34dfdrp+s~4VEM)]euVEM2\displaystyle-f(r_{0})\left[(\beta+1)f\left(\frac{3}{4}\frac{df}{dr_{p}}+\frac{\tilde{s}}{4}V_{EM}\right)e^{v^{\prime}\frac{V_{EM}}{2}}\right]_{r=r_{0}}+\int_{t-|x|}^{t^{\prime\prime}}du^{\prime}\,\frac{d}{dr_{p}}\left[(\beta-1)f\left(\frac{3}{4}\frac{df}{dr_{p}}+\frac{\tilde{s}}{4}V_{EM}\right)\right]e^{u^{\prime}\frac{V_{EM}}{2}}
+t′′x(r0)𝑑vddrp[(β+1)f(34dfdrp+s~4VEM)]evVEM2,\displaystyle+\int_{t^{\prime\prime}}^{x(r_{0})}dv^{\prime}\,\frac{d}{dr_{p}}\left[(\beta+1)f\left(\frac{3}{4}\frac{df}{dr_{p}}+\frac{\tilde{s}}{4}V_{EM}\right)\right]e^{v^{\prime}\frac{V_{EM}}{2}},

where δt′′δr0=r03/22(r02)η\frac{\delta t^{\prime\prime}}{\delta r_{0}}=\frac{r_{0}^{3/2}}{2(r_{0}-2)}\eta. Hence, the EM perturbation is

Φlmdip(1)e(r,t)\displaystyle{}_{e}\Phi^{dip(1)}_{lm}(r,t) =qλYlmηΦ^dip(1)(r,t),\displaystyle=-\frac{q}{\lambda}Y_{lm}^{*}\eta\,\hat{\Phi}^{dip(1)}(r,t), (63)

where ++ and - refer to positive and negative charges respectively, and p=qηp=q\eta, assuming that the scale of change of rpzeror_{pzero} is much larger than pp. In a region outside the dipole, there is no monopole contribution. The radiative contribution is represented by

Frade1=p(0[δ2(θ,θ0,ϕ,ϕ0)14π]Φ^dip(1)(t,r)r20GS2θfΦ^dip(1)(t,r)rGS2θ1fΦ^dip(1)(t,r)t00GS2ϕfΦ^dip(1)(t,r)rGS2ϕ1fΦ^dip(1)(t,r)t00),\begin{aligned} {}_{e1}F_{\text{rad}}&=-p\begin{pmatrix}0&**&**&**\\ \left[\delta^{2}(\theta,\theta_{0},\phi,\phi_{0})-\frac{1}{4\pi}\right]\frac{\hat{\Phi}^{dip(1)}(t,r)}{r^{2}}&0&**&**\\ -\frac{\partial G_{S^{2}}}{\partial\theta}f\frac{\partial\hat{\Phi}^{dip(1)}(t,r)}{\partial r}&-\frac{\partial G_{S^{2}}}{\partial\theta}\frac{1}{f}\frac{\partial\hat{\Phi}^{dip(1)}(t,r)}{\partial t}&0&0\\ -\frac{\partial G_{S^{2}}}{\partial\phi}f\frac{\partial\hat{\Phi}^{dip(1)}(t,r)}{\partial r}&-\frac{\partial G_{S^{2}}}{\partial\phi}\frac{1}{f}\frac{\partial\hat{\Phi}^{dip(1)}(t,r)}{\partial t}&0&0\end{pmatrix}\end{aligned}, (64)

where δ2(θ,θ0,ϕ,ϕ0)\delta^{2}(\theta,\theta_{0},\phi,\phi_{0}) is the Dirac delta function at the radial line of free fall, while GS2G_{S^{2}} is the angular Green’s function on a S2S^{2} sphere (the expressions can be found in [40]). There is also another contribution to the electromagnetic field, localized at the position of the points charges, which is referred as Finte2{}_{e2}F_{\text{int}}, given by

Finte2\displaystyle{}_{e2}F_{\text{int}} =p(0000GS2θf[δ(rrp(t))δ(rrp(t)η)]GS2θβ[δ(rrp(t))δ(rrp(t)η)]00GS2ϕf[δ(rrp(t))δ(rrp(t)η)]GS2ϕβ[δ(rrp(t))δ(rrp(t)η)]00).\displaystyle=-p\begin{pmatrix}0&0&**&**\\ 0&0&**&**\\ -\frac{\partial G_{S^{2}}}{\partial\theta}f\left[\delta(r-r_{p}(t))-\delta(r-r_{p}(t)-\eta)\right]&-\frac{\partial G_{S^{2}}}{\partial\theta}\beta\,\left[\delta(r-r_{p}(t))-\delta(r-r_{p}(t)-\eta)\right]&0&0\\ -\frac{\partial G_{S^{2}}}{\partial\phi}f\left[\delta(r-r_{p}(t))-\delta(r-r_{p}(t)-\eta)\right]&-\frac{\partial G_{S^{2}}}{\partial\phi}\beta\,\left[\delta(r-r_{p}(t))-\delta(r-r_{p}(t)-\eta)\right]&0&0\end{pmatrix}. (65)

Similar to FmonopoleF_{\text{monopole}}, the term Finte2{}_{e2}F_{\text{int}} vanishes outside the dipole. Consequently, its contribution is ignored when calculating the electromagnetic stress-energy tensor TEMμνT_{EM}^{\mu\nu}, which will source the gravitational perturbation.

VI Numerical solution

In this section, we discuss the results of numerically integrating the contributions to the QNM and flat electromagnetic perturbations arising from a radially free-falling radial dipole. In the figures, the particle’s position is marked by a red dashed vertical line, while the peak of the potential, VEM=M10V_{\text{EM}}=\frac{M}{10}, is also indicated. The dipole initially begins around x0=200Mx_{0}=200M with a separation η=1M\eta=1M.

For the QNM, initially at t=0Mt=0M, in Figure 3(a), there is no contribution to Φ^QNM(1)dip\hat{\Phi}^{(1)\text{dip}}_{\text{QNM}}. As time progresses, a constant pulse emerges (see Figure 3(b)), propagating symmetrically from the peak of the potential and moving to the left and right. This event occurs when the particle reaches xp=175.5Mx_{p}=175.5M at t=231.25Mt=231.25M. This pulse might be related to the fact that charge qq enters the light cone of (0,tx)(0,t-x), while the other charge, for a short period, remains outside. This highlights two artifacts in this problem: the modeling of the dipole as two discrete charges, and the sharp causality boundaries within the QNM region due to the Dirac delta potential approximation. Additionally, reducing η\eta would control the width of this pulse, while a more realistic, extended potential might reduce its amplitude.

Later, as the particle crosses the potential peak at x=0x=0, as shown in (e), around t=1282Mt=1282M, part of ΦQNM(1)dip\Phi^{(1)\text{dip}}_{\text{QNM}} reflects, while another part transmits with nearly equal amplitudes. The transmitted component continues toward the horizon, following the particle’s trajectory, as shown in (f) and (g). When analyzed on a logarithmic scale, the slope of log[ΦQNM(1)dip]\log\left[\Phi^{(1)\text{dip}}_{\text{QNM}}\right] can be approximately fitted with a line of slope VEM2-\frac{V_{\text{EM}}}{2}, suggesting a QNM excitation with an almost constant amplitude, consistent with our expectations.

Around t=1596Mt=1596M, the particle is nearly frozen at r=2Mr=2M, and all QNM contributions to the perturbation vanish completely, as also shown in (h). After some time, around t=1630Mt=1630M and t=1670Mt=1670M, two additional pulses and a shock wave are excited at x=0x=0 and begin propagating leftward and rightward, as shown in (i). The second shock wave might suggest that another QNM is excited at late times.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Refer to caption
(e)
Refer to caption
(f)
Refer to caption
(g)
Refer to caption
(h)
Refer to caption
(i)
Figure 3: Visual representation of different snapshots of time of the QNM part of the perturbation ΦQNM(1)dip\Phi^{(1)\text{dip}}_{\text{QNM}}. The xx-range is from 200-200 to 200M200M, while tt ranges from 0 to 1850M1850M.

For the flat part, as shown in Figure 4(a), (b), and (c), there is a residual field contribution that does not decay and remains constant across the entire spacetime. The field flips sign across the particle, which is expected for a dipole moment. However, this nonlocal behavior is unrealistic, attributable to the fact that we are solving the wave equation in a 1D variable xx, while replacing the entire potential with a Dirac delta. This excludes the centrifugal effects, which represent the impact of the other two coordinates. Consequently, if a better approximation were made that preserved the Minkowski limit, we would expect this field to be local and to decay as r3r^{-3} away from the potential, as a dipole would in flat spacetime. This suggests that a more accurate approximation could be adopted in future studies if we aim to explore the perturbation more thoroughly.

Furthermore, as the particle approaches the potential, the field builds residual strength near the potential, suggesting that some part of the field will be reflected due to the barrier, as shown in (d) and (e). In (f), we observe a portion of this field being reflected. We anticipate that in a more realistic scenario, as the source of the electromagnetic field moves closer to the potential peak, some of its local (non-radiative) fields would begin to decouple and radiate away, as shown in (g) in our simplified model. Once this shockwave has traveled far enough, all flat contributions vanish, as shown in (i).

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Refer to caption
(e)
Refer to caption
(f)
Refer to caption
(g)
Refer to caption
(h)
Refer to caption
(i)
Figure 4: Visual of different snapshots of time of the flat part of the perturbation ΦFlat(1)dip\Phi^{(1)\text{dip}}_{\text{Flat}}. The xx-range is from 200-200 to 200M200M, while tt ranges from 0 to 1850M1850M.

VII Discussion & Conclusion

In this work, we have studied the electromagnetic perturbations at first Φ(1)\Phi^{(1)} and second order Φ(2)\Phi^{(2)}, using the Regge-Wheeler formalism in the Schwarzschild spacetime. We began by briefly summarizing relevant literature on the first-order electromagnetic perturbations in Schwarzschild coordinates, then reformulated the Faraday tensor FμνF_{\mu\nu} and electromagnetic stress-energy tensor TμνT_{\mu\nu} in terms of the even and odd perturbation scalars Φlm(1)o,e{}_{o,e}\Phi^{(1)}_{lm}, as well as the current multipoles j(i)lmj_{(i)\,lm}.

For the second-order perturbations, we derived the effective source terms for the second-order Regge-Wheeler master equation (with s=1s=1). These sources are quadratic in the first-order gravitational perturbations (both even and odd), given by the Zerilli and Regge-Wheeler master equations (with s=2s=2), and the electromagnetic perturbations (both odd and even). We accounted only for the electromagnetic current JJ in these sources, which contributes to induced modes that are linear in gravitational perturbations with an amplitude dependent on the details of the electromagnetic potential. However, for brevity, we did not consider the perturber’s effective stress-energy source term TGμνT^{\mu\nu}_{G} in the gravitational perturbations, which could be of some relevance for EMRI studies. We note that including these terms would induce a part of the source in the second-order electromagnetic perturbation that does not depend on first-order perturbations but rather on the physical parameters of the perturber, such as its mass MM and charge qq, along with a linear electromagnetic perturbation whose amplitude depends on both electromagnetic and gravitational details.

Furthermore, by replacing the Regge-Wheeler potential with a Dirac delta function as well as considering a sample source similar to those used in [1, 35], and representing a product of linear gravitational and linear electromagnetic perturbations of frequencies ω(1)=iVG2\omega^{(1)}=-i\frac{V_{G}}{2} and Ω(1)=iVEM2\Omega^{(1)}=-i\frac{V_{EM}}{2} respectively, we illustrated the mixing that takes place at the second-order electromagnetic QNM ΦQNM(2)\Phi^{(2)}_{QNM} with a frequency of Ω(2)=iVEM+VG2\Omega^{(2)}=-i\frac{V_{EM}+V_{G}}{2}.

Studying the mixing of the second-order electromagnetic perturbations complements our study of the mixing of second-order gravitational perturbations in [1]. We showed that, when examining the minimally coupled Einstein-Maxwell field equations perturbatively, second-order corrections are induced that are quadratic in the electromagnetic perturbations gravity sector and quadratic in the electromagnetic and gravitational perturbations for electromagnetic sector. Besides its theoretical significance, this mixing could be relevant for multi-messenger astrophysics, particularly for stellar mass and EMRI mergers observed by Advanced LIGO and the upcoming LISA. For more discussion on that we refer the reader to [1].

Moreover, we noticed that changing the coupling details, for example by considering the Maxwell-Dilaton-Einstein equations, induces a different mixing spectrum. In this context, this mixing phenomenon is relevant not only for testing the minimal coupling principle but also for probing modified theories of gravity.

For the first-order perturbation, we reduced the calculations of the perturbation from a point charge source to a one-dimensional path integral. A suitable choice of parameterization can simplify some of these integrals. By approximating the Regge-Wheeler potential as a Dirac delta function and focusing on two cases — a single charge qq and a radial dipole moment p=qηp=q\eta — we found that the flat part of the perturbation is easy to derive analytically, though the QNM part requires a semi-analytical approximation or numerical implementation. We carried out both methods in this work. For the dipole case, based on the numerical solution, we concluded that the QNM perturbation is excited with an almost constant amplitude.

Nevertheless, the Dirac delta function approximation presents some limitations, as discussed in the numerical integration section. By approximating the full potential with a Dirac delta, we neglect the centrifugal correction, which encapsulates the Minkowski limit of the field’s behavior and effectively reduces the problem to one dimension. In this 1D context, the field does not decay as a function of rr, which is unrealistic for a 3D problem. This same issue leads to singular multipole coefficients in the stress-energy tensor reported in Appendix A of [1]. On the other hand, a less trivial approximation would be harder to handle analytically and would not yield simple modes like those with the Dirac delta function. Thus, the choice of approximation should depend on the specific aspects intended for investigation. In our case, as we are interested in mixing, a more realistic potential would be beneficial, and in future work we intend to revisit the mixing problem in Teukolsky equations in the Kerr background, using numerical approach.

Acknowledgements.
We acknowledge the partial support of D.S. through the U.S. National Science Foundation under Grant No. PHY-2310363. Additionally, we express our gratitude to Macarena Lagos, Ariadna Ribes Metidieri, Ahmed Elshahawy, A.K. Gorbatsievich, and Stanislav Komarov for their insightful discussions. Special thanks to Jacob Fields and Ish Gupta for their invaluable guidance and feedback on this manuscript. We also sincerely appreciate the valuable comments provided by David Radice and Elias Most. Finally, we extend special thanks to Mahmoud A. Mansour for his invaluable time, support, and assistance throughout this project.

Appendix A Green’s function on S2S^{2}

The Green’s function GS2G_{S^{2}} on S2S^{2} with a unit radius is

S2GS2\displaystyle\nabla_{S^{2}}G_{S^{2}} =Y00(θ,ϕ)Y00(θ0,ϕ0)+δ(θθ0)δ(ϕϕ0)sinθ\displaystyle=-Y_{00}(\theta,\phi)Y_{00}^{*}(\theta_{0},\phi_{0})+\frac{\delta(\theta-\theta_{0})\delta(\phi-\phi_{0})}{\sin\theta} (66)
=l>0,mYlm(θ,ϕ)Ylm(θ0,ϕ0),\displaystyle=\sum_{l>0,m}Y_{lm}(\theta,\phi)Y_{lm}^{*}(\theta_{0},\phi_{0}),
GS2\displaystyle G_{S^{2}} =l>0,m1λYlm(θ,ϕ)Ylm(θ0,ϕ0)\displaystyle=\sum_{l>0,m}\frac{-1}{\lambda}Y_{lm}(\theta,\phi)Y_{lm}^{*}(\theta_{0},\phi_{0})
=14πln[1cosγ]\displaystyle=\frac{1}{4\pi}\ln\left[1-\cos\gamma\right]

where cosγ=cosθcosθ0+sinθsinθ0cos(ϕϕ0),\cos\gamma=\cos\theta\cos\theta_{0}+\sin\theta\sin\theta_{0}\cos(\phi-\phi_{0}), and γ\gamma is the angle between the two points (θ,ϕ)(\theta,\phi) and (θ0,ϕ0)(\theta_{0},\phi_{0}).

GS2ϕ=sinθsinθ0sin(ϕϕ0)4π(1cosγ)\frac{\partial G_{S^{2}}}{\partial\phi}=-\frac{\sin\theta\sin\theta_{0}\sin(\phi-\phi_{0})}{4\pi(1-\cos\gamma)}
GS2θ=cosθ0sinθ+cosθcos(ϕϕ0)sinθ04π(1cosγ)\frac{G_{S^{2}}}{\partial\theta}=\frac{-\cos\theta_{0}\sin\theta+\cos\theta\cos(\phi-\phi_{0})\sin\theta_{0}}{4\pi(1-\cos\gamma)}

Appendix B Integrals

Here, ϵrpBzero(t,r)\epsilon_{r_{pBzero}(t,r)} is a binary indicator that equals 1 if BB has a zero at a given (t,r)(t,r), and 0 otherwise. The quantities rpAmin(t,r)r_{pA_{\min}}(t,r) and rpBmin(t,r)r_{pB_{\min}}(t,r) denote the minimal positive values of AA and BB, respectively, while rpAmax(t,r)r_{pA_{\max}}(t,r) and rpBmax(t,r)r_{pB_{\max}}(t,r) represent their corresponding maximal positive values.

For a given (t,r)(t,r), if A=0A=0 (or B=0B=0), it follows that rpAmin=0r_{pA_{\min}}=0 (or rpBmin=0r_{pB_{\min}}=0). Both AA and BB are monotonic functions with respect to the parameter rpr_{p} along the particle’s trajectory for a fixed (t,r)(t,r) point. The partial derivatives of AA and BB with respect to rpr_{p} are positive definite

Arp=sβ(rp)2+1β(rp)sβ(rp)+1>0\frac{\partial A}{\partial r_{p}}=s\beta(r_{p})^{2}+\frac{1}{\beta(r_{p})}-\frac{s}{\beta(r_{p})+1}>0 (67)
Brp=s~f(rp)+1β(rp)+β(rp)f(rp)[1s~β(rp)]>0\frac{\partial B}{\partial r_{p}}=-\tilde{s}f(r_{p})+\frac{1}{\beta(r_{p})}+\frac{\beta(r_{p})}{f(r_{p})}\left[1-\tilde{s}\beta(r_{p})\right]>0 (68)

Consequently, the causal structure of the Green’s function confines the particle’s support to a (t,r)(t,r) cone along its trajectory. In summary, the problem reduces to finding max, min and zero (if it exists) for A=0A=0 and B=0B=0 as a function of (t,r)(t,r), as well as evaluating the two integrals IQ1BI_{Q1B} and IQ2BI_{Q2B} which resemble part of the contribution of the QNM mode.

IF1A=2M𝑑rpf(rp)β(rp)df(rp)drpΘ(A)=[23β32β]|rpAmin(t,r)rpAmax(t,r),\displaystyle I_{F1A}=\int_{2M}^{\infty}dr_{p}\frac{f\left(r_{p}\right)}{\beta\left(r_{p}\right)}\frac{df\left(r_{p}\right)}{dr_{p}}\Theta(A)=\left[\frac{2}{3}\beta^{3}-2\beta\right]\bigg{|}_{r_{pA_{\min}}(t,r)}^{r_{pA_{\max}}(t,r)}, (69)
IF1B=2M𝑑rpf(rp)β(rp)df(rp)drpΘ(B)=[23β32β]|rpBmin(t,r)rpBmax(t,r),\displaystyle I_{F1B}=\int_{2M}^{\infty}dr_{p}\frac{f\left(r_{p}\right)}{\beta\left(r_{p}\right)}\frac{df\left(r_{p}\right)}{dr_{p}}\Theta(B)=\left[\frac{2}{3}\beta^{3}-2\beta\right]\bigg{|}_{r_{pB_{\min}}(t,r)}^{r_{pB_{\max}}(t,r)}, (70)
IQ1B=2M𝑑rpf(rp)β(rp)df(rp)drpeVEM2(t(rp)+|x(rp)|)Θ(B)=rpBmin(t,r)rpBmax(t,r)𝑑rpf(rp)β(rp)df(rp)drpeVEM2(t(rp)+|x(rp)|),\displaystyle I_{Q1B}=\int_{2M}^{\infty}dr_{p}\,\frac{f(r_{p})}{\beta(r_{p})}\frac{df(r_{p})}{dr_{p}}\,e^{\frac{V_{EM}}{2}\left(t(r_{p})+\left|x(r_{p})\right|\right)}\Theta(B)=\int_{r_{pB_{\min}}(t,r)}^{r_{pB_{\max}}(t,r)}dr_{p}\,\frac{f(r_{p})}{\beta(r_{p})}\frac{df(r_{p})}{dr_{p}}\,e^{\frac{V_{EM}}{2}\left(t(r_{p})+\left|x(r_{p})\right|\right)}, (71)
IF2A=2M𝑑rpf(rp)β(rp)δ(A)=ϵrpAzero(t,r)f(rpAzero(t,r))β(rpAzero(t,r))I_{F2A}=\int_{2M}^{\infty}dr_{p}\frac{f\left(r_{p}\right)}{\beta\left(r_{p}\right)}\delta(A)=\epsilon_{r_{pAzero}(t,r)}\frac{f\left(r_{pAzero}(t,r)\right)}{\beta\left(r_{pAzero}(t,r)\right)} (72)
IQ2B=2M𝑑rpf(rp)β(rp)eVEM2(t(rp)+|x(rp)|)Θ(B)=rpBmin(t,r)rpBmax(t,r)𝑑rpf(rp)β(rp)eVEM2(t(rp)+|x(rp)|),I_{Q2B}=\int_{2M}^{\infty}dr_{p}\frac{f\left(r_{p}\right)}{\beta\left(r_{p}\right)}e^{\frac{V_{EM}}{2}\left(t\left(r_{p}\right)+\left|x\left(r_{p}\right)\right|\right)}\Theta(B)=\int_{r_{pB_{\min}}(t,r)}^{r_{pB_{\max}}(t,r)}dr_{p}\frac{f(r_{p})}{\beta(r_{p})}\,e^{\frac{V_{EM}}{2}\left(t(r_{p})+\left|x(r_{p})\right|\right)}, (73)

Appendix C Re-parameterization of perturbation integral equation

We can parameterize the path integral in terms of rpr_{p} as well:

Φlmo(r,t)=\displaystyle{}_{o}\Phi_{lm}(r^{\prime},t^{\prime})= qλ2Mdrpf(rp){imdθ^drp𝒞lm[Ylm]\displaystyle-\frac{q}{\lambda}\int_{2M}^{\infty}dr_{p}\,f\left(r_{p}\right)\left\{im\frac{d\hat{\theta}}{dr_{p}}\mathcal{C}_{lm}[Y_{lm}^{*}]\right. (74)
+dϕ^drp𝒟lm[Ylm]}Glm(r,t,rp,t(rp)).\displaystyle\left.+\frac{d\hat{\phi}}{dr_{p}}\mathcal{D}_{lm}[Y_{lm}^{*}]\right\}\,G_{lm}\left(r^{\prime},t^{\prime},r_{p},t\left(r_{p}\right)\right).
e Φlm(t,r)=qλ2Mdrp{[dr˙p(rp)drpf(rp)r˙p(rp)df(rp)drp\displaystyle\Phi_{lm}\left(t^{\prime},r^{\prime}\right)=\frac{q}{\lambda}\int^{\infty}_{2M}dr_{p}\Bigg{\{}\left[\frac{d\dot{r}_{p}(r_{p})}{dr_{p}}-\frac{f\left(r_{p}\right)}{\dot{r}_{p}\left(r_{p}\right)}\frac{df\left(r_{p}\right)}{dr_{p}}\right. (75)
imr˙(rp)dϕ^drp+mcotθ^r˙(rp)dθ^drp]Ylm\displaystyle\left.-im\,\dot{r}\left(r_{p}\right)\frac{d\hat{\phi}}{dr_{p}}+m\cot\hat{\theta}\,\dot{r}\left(r_{p}\right)\frac{d\hat{\theta}}{dr_{p}}\right]Y_{lm}^{*}
+[(lm)(l+m+1)eiϕ^r˙(rp)dθ^drp]Yl,m+1}\displaystyle+\left[\sqrt{(l-m)(l+m+1)}e^{i\hat{\phi}}\dot{r}\left(r_{p}\right)\frac{d\hat{\theta}}{dr_{p}}\right]Y_{l,m+1}^{*}\Bigg{\}}
×Glm(r,t,rp,t(rp))\displaystyle\times G_{lm}\left(r^{\prime},t^{\prime},r_{p},t\left(r_{p}\right)\right)
+qλ2M𝑑rpGlmdr|r=rp[r˙pf2(rp)r˙p]Ylm.\displaystyle+\frac{q}{\lambda}\int^{\infty}_{2M}dr_{p}\left.\frac{G_{lm}}{dr}\right|_{r=r_{p}}\left[\dot{r}_{p}-\frac{f^{2}\left(r_{p}\right)}{\dot{r}_{p}}\right]Y_{lm}^{*}.

where ϕ^(rp)=ϕ(t(rp))\hat{\phi}(r_{p})=\phi(t(r_{p})) and θ^(rp)=θ(t(rp))\hat{\theta}(r_{p})=\theta(t(r_{p})). This re-parameterization in terms of rpr_{p} is valid as long as the particle undergoes radial motion in Schwarzschild coordinates. Still, if the particle passes through the same radial point at different moments in time, the integral needs to be divided into the one-to-on domains and be handled separately. On the other hand, for a purely circular trajectory, we should re-parameterize using the angular coordinate instead.

Appendix D 4D vector harmonics

In this appendix, spherical decomposition in [23] is extended to 4D vector representations, following the same approach of Lousto and Barack [41], but for vectors instead of tensors. The harmonics are defined on the two-dimensional sphere S2S^{2}, where scalar spherical harmonics Ylm(θ,ϕ)Y_{lm}(\theta,\phi) orthogonality are extended to the 4D vector harmonics 𝐘μlm(θ,ϕ)\mathbf{Y}_{\mu\,lm}(\theta,\phi). The vector harmonics Y(1)lmμ,Y(2)lmμ,Y(3)lmμY^{\mu}_{(1)\,lm},Y^{\mu}_{(2)\,lm},Y^{\mu}_{(3)\,lm} are categorized as even-parity, while Y(4)lmμY^{\mu}_{(4)\,lm} belongs to the odd-parity class. These are expressed as:

𝐘(1)lmμ=(Ylm000),𝐘(2)lmμ=(0Ylm00),\mathbf{Y}^{\mu}_{(1)\,lm}=\left(\begin{array}[]{c}Y_{lm}\\ 0\\ 0\\ 0\end{array}\right),\quad\mathbf{Y}^{\mu}_{(2)\,lm}=\left(\begin{array}[]{c}0\\ Y_{lm}\\ 0\\ 0\end{array}\right), (76)
𝐘(3)lmμ=1λ(00Y˙lmcsc2θYlm),𝐘(4)lmμ=1λ(00cscθYlmcscθY˙lm).\mathbf{Y}^{\mu}_{(3)\,lm}=\frac{1}{\lambda}\left(\begin{array}[]{c}0\\ 0\\ \dot{Y}_{lm}\\ \csc^{2}\theta Y_{lm}^{\prime}\end{array}\right),\quad\mathbf{Y}^{\mu}_{(4)\,lm}=\frac{1}{\lambda}\left(\begin{array}[]{c}0\\ 0\\ \csc\theta Y_{lm}^{\prime}\\ -\csc\theta\dot{Y}_{lm}\end{array}\right). (77)

where Y˙lm\dot{Y}_{lm} and YlmY_{lm}^{\prime} represent the derivatives with respect to θ\theta and ϕ\phi, respectively. Their is ll and mm sub-index label on the vector harmonics 𝐘(i)lmμ\mathbf{Y}^{\mu}_{(i)\,lm}, but suppressed for simplicity. Using the following two identities:

S2𝑑Ωcscθ(YlmY˙lmYlmY˙lm)=0,\int_{S^{2}}d\Omega\csc\theta\,(Y^{*\prime}_{lm}\,\dot{Y}_{lm}-Y^{\prime}_{lm}\,\dot{Y}^{*}_{lm})=0, (78)

and

S2𝑑Ω(csc2θYlmYlm+Y˙lmY˙lm)=δllδmmλ.\int_{S^{2}}d\Omega(\csc^{2}\theta\,Y^{\prime}_{lm}\,Y^{*\prime}_{lm}+\dot{Y}_{lm}\,\dot{Y}^{*}_{lm})=\delta_{ll^{\prime}}\delta_{mm^{\prime}}\lambda. (79)

The vector harmonics orthogonality condition is given by:

S2𝐘(i)lmμ𝐘μ(j)lm𝑑Ω=δijδllδmmNi(r),\int_{S^{2}}\mathbf{Y}^{\mu}_{(i)lm}\mathbf{Y}_{\mu(j)l^{\prime}m^{\prime}}\,d\Omega=\delta_{ij}\delta_{ll^{\prime}}\delta_{mm^{\prime}}N_{i}(r), (80)

where Ni(r)\sqrt{N_{i}(r)} represents the norm of the ii-th vector harmonic. The components of the even-parity vectors 𝐉e{}_{e}\mathbf{J} and 𝐀e{}_{e}\mathbf{A} are expressed as:

𝐉μe=(j1Yj2Yj3Y˙j3Y)𝐀μe=(gYhYkY˙kY){}_{e}\mathbf{J}_{\mu}=\begin{pmatrix}j_{1}Y\\ j_{2}Y\\ j_{3}\dot{Y}\\ j_{3}Y^{\prime}\end{pmatrix}\quad{}_{e}\mathbf{A}_{\mu}=\begin{pmatrix}gY\\ hY\\ k\dot{Y}\\ kY^{\prime}\end{pmatrix} (81)

Similarly, the components of the odd-parity vectors 𝐀μo{}_{o}\mathbf{A}_{\mu} and 𝐉μo{}_{o}\mathbf{J}_{\mu} are given by:

𝐉μo=(00j4cscθYj4sinθY˙)𝐀μo=(00acscθYasinθY˙.){}_{o}\mathbf{J}_{\mu}=\begin{pmatrix}0\\ 0\\ j_{4}\csc\theta Y^{\prime}\\ -j_{4}\sin\theta\,\dot{Y}\end{pmatrix}\quad{}_{o}\mathbf{A}_{\mu}=\begin{pmatrix}0\\ 0\\ a\csc\theta Y^{\prime}\\ -a\sin\theta\,\dot{Y}.\end{pmatrix} (82)

The summation over the angular ll and azimuthal mm numbers is implied but not explicitly shown. The functions ji,a,g,kj_{i},a,g,k are dependent on the coordinates (t,r)(t,r) and the sub-index lmlm.

References