• No results found

Geometrical and Perturbative study of Tidally Deformed Schwarzschild Spacetime

N/A
N/A
Protected

Academic year: 2021

Share "Geometrical and Perturbative study of Tidally Deformed Schwarzschild Spacetime"

Copied!
64
0
0

Loading.... (view fulltext now)

Full text

(1)

Master Thesis

Geometrical and Perturbative study of Tidally Deformed Schwarzschild

Spacetime

Division of Theoretical Physics Department of Physics and Astronomy

Uppsala University

Author:

Zeyd Sam

Supervisor:

Asst.Prof. Giuseppe Dibitetto Subject Reader:

Prof. Ulf Danielsson

(2)

Abstract

We undertake a study of the tidal deformation of a Schwarzschild black hole. We formulate the geometry in terms of intrinsic and extrinsic quantities on the event horizon, which is described by a null hypersurface embedded in spacetime, and we do so by formulating the Gauss-Codazzi theory of null hypersurfaces, extending standard results for the spacelike and timelike cases. This formalism is then applied to a solution of the Einstein field equations in vaccum, in order to describe tidal distortions to the black hole horizon due to a small orbiting body. The techniques that we employ are valid for rapid orbits in the strong field of the black hole. This analysis relies on a perturbative approach for bodies with a very large mass ratio. We conduct the study in the frequency domain of black hole perturbation theory, for a small body orbiting in a circular orbit. The results are visualized by embedding the distorted horizon in Euclidean space, showing how the horizon is deformed from an embed- ded sphere to an ellipsoid when the small body is close to the black hole, and when the orbital separation is increased, showing in both cases the angular offset between the horizon bulge and the body that induces it.

For both orbital separations considered, we provide snapshots of the tidal distortion as the small body orbits around the black hole.

Popul¨arvetenskaplig sammanfattning

Vi unders¨oker tidvattendeformationen av ett icke-roterande, s˚a kallade Schwarzschild, svarta h˚al. Vi formulerar geometrin i form av inre och yttre geometriska kvantitet p˚a h¨andelsehorisonten, som beskrivs av en hyperyta inb¨addad i rumtiden, och vi g¨or det genom att formulera Gauss- Codazzi teorin om nollhyperytor som str¨acker det standard resultat f¨or rumslika och tidslika fallen. Denna formalism till¨ampas sedan p˚a en l¨osning av Einstein-f¨altekvationerna i vaccum, f¨or att beskriva tidvat- tensf¨orvr¨angningar till svarth˚alshorisonten p˚a grund av en liten kretsande objekt. De tekniker som vi anv¨ander ¨ar giltiga f¨or sm˚a objekt som r¨or sig snabbt i omloppsbanor, i det svarta h˚alets starka gravitationsf¨alt. Denna analys bygger p˚a en perturbativ strategi f¨or tv˚a kroppar med ett mycket stort massf¨orh˚allande - det vill s¨aga, det sm˚a objektet ¨ar mycket mindre

¨

an den andra. Vi genomf¨or studien i frekvensdom¨anen f¨or st¨orningsteori f¨or svart h˚al, f¨or en liten kropp som g˚ar i en cirkul¨ar bana. Resultaten visualiseras genom att b¨adda in den f¨orvr¨angda horisonten i det euklidiska rymden, vilket visar hur horisonten deformeras fr˚an en inb¨addad sf¨ar till en ellipsoid n¨ar den lilla kroppen ¨ar n¨ara det svarta h˚alet, och n¨ar orbita- lavskiljningen ¨okar, vilket i b˚ada fallen visar vinkelf¨orskjutning mellan ho- risontbukten och kroppen som producerar den. F¨or b˚ada orbitalskillnader som beaktas tillhandah˚aller vi ¨ogonblicksbilder av tidvattenf¨orvr¨angnin- gen n¨ar den mindre kroppen kretsar runt det svarta h˚alet.

Uploaded Oct 20 for access, Uppsala universitet.

(3)

Contents

1 Introduction 4

1.1 Context and motivation . . . 4

1.2 Two-body problem in GR . . . 4

1.3 Tidal distortion of Newtonian fluid body . . . 6

1.4 Tidal distortion of compact objects . . . 7

1.5 Outline of project. . . 8

2 Hypersurface geometry 10 2.1 Constructing a null hypersurface . . . 10

2.2 Derivation of Gauss-Weingarten equations . . . 12

2.3 Intrinsic and extrinsic quantities . . . 13

2.4 Derivation of Gauss-Codazzi equations . . . 14

2.5 Components of Ricci tensor . . . 16

2.6 Evolution of expansion and shear . . . 19

3 Perturbed Schwarzschild black hole 22 3.1 Schwarzschild spacetime . . . 22

3.2 Metric perturbation . . . 23

3.3 Horizon distortion . . . 23

3.4 First horizon equation . . . 25

3.5 Second set of horizon equations . . . 27

3.6 Intrinsic horizon geometry . . . 29

3.7 Extrinsic horizon geometry . . . 30

4 Black hole perturbation theory 32 4.1 Even-parity master function . . . 32

4.2 Odd-parity master function . . . 33

4.3 Near-horizon analysis. . . 33

4.4 Measures of tidal distortion . . . 34

4.5 Computing the Teukolksy amplitudes . . . 35

4.6 Horizon curvature from Teukolsky amplitudes . . . 37

5 Orbits about Schwarzschild 38 5.1 Orbital energy . . . 38

5.2 Embedding of horizon in Euclidean space . . . 39

5.3 Dephasing of tidal distortion . . . 41

6 Numerical analysis 42 6.1 Teukolsky code . . . 42

6.2 Mode convergence . . . 44

6.3 Visualizing a deformed Schwarzschild horizon . . . 46

(4)

7 Discussion and conclusions 49 7.1 Summary . . . 49 7.2 Further directions. . . 51 7.3 Conclusion . . . 53

A Reparameterization of null generators 55

B Spherical harmonics 56

B.1 Scalar spherical harmonics . . . 56 B.2 Vector spherical harmonics . . . 57 B.3 Tensor spherical harmonics . . . 58

C Calculation of induced metric 59

D Calculation on extrinsic horizon 60

D.1 Perturbative calculation of connection coefficients . . . 61

References 62

(5)

1 Introduction

1.1 Context and motivation

The astounding first detection of gravitational waves by LIGO (Laser Interfer- ometer Gravitational Wave Observatory) in 2015 [1] not only validated one of the famed predictions of Einstein’s theory of gravitation, the celebrated theory of General Relativity, but it also ushered in a new era in how we observe our Uni- verse - gravitational wave astronomy. No longer are we bound just by ”sight”, but are we are now attuned to the ”sound” of spectacular events unfolding in the cosmos. This first detection came from the orbiting inspiral and merger of two black holes, of 36 and 29 solar masses, and was awarded the 2017 Nobel prize in physics. Unlike orbits in the Newtonian theory of gravity, accelerating bodies in GR produce gravitational waves, which manifest as ripples in the fabric of spacetime, carrying energy and angular momentum and traveling radially away from the source at the speed of light. This loss of orbital energy and angular momentum due to gravitational wave emission causes the bodies to spiral into one another and merge. The two-body problem, as it is known, is thus far more complicated in GR than in Newtonian theory, and the key to understanding the rich information carried by gravitational waves from compact binaries lies in accurately modeling the two-body problem for a range of masses, spins, and orbital configurations. A lot of work is being done on this front, for detection with current and future gravitational wave detectors.

1.2 Two-body problem in GR

As mentioned, a detailed understanding of the dynamics of inspiraling com- pact objects (black holes and neutron stars) is required in order to understand the information carried by gravitational waves. So how do we go about this?

There exists a collection of exact solutions to the Einstein field equations [2].

However, apart from a handful of solutions, such as the trivial Minkowski space- time, Schwarzschild Reissner-Nordstr¨om and Kerr black hole solutions, and the Friedmann-Robertson-Walker spacetime encountered in cosmology along with de Sitter and anti-de Sitter spacetimes [3], most of the solutions are algebraic in value and offer little to no physical interpretation. From the physical black hole solutions that we have listed, none can do the job of accurately describing and evolving the dynamics of a pair of coalescing compact objects and the particular gravitational waveform they produce, for a set of initial conditions.

So how does one go about solving the two-body problem in GR? Given a set of initial conditions, we wish to evolve the system forward in time. To do this, we must solve the Einstein field equations (EFE) for this complex system.

This is easier said than done. The EFE are a complicated set of ten coupled and highly nonlinear partial differential equations given by

Gαβ= 8πG

c4 Tαβ (1.1)

(6)

where Gαβ is the Einstein tensor, Tαβ is the energy-momentum tensor of the system, G is the universal gravitational constant and c the velocity of light in vacuum. Often one works in natural units in which G = c = 1. Can we get an exact solution to this problem, or do we have to approximate it? We follow [4]

to give an overview.

One method is to use numerical relativity. Using powerful supercomputers and sophisticated computational techniques, one can solve the EFE numerically and propagate the solution forward in time. We can regard numerical relativity as being an almost exact method of solution (if we ignore inevitable numerical errors arising from the computations). This technique is well suited to compact bodies of comparable mass, such as the first event detected by LIGO. In their last moments before merger, the two bodies are orbiting very closely about one another, and at a significant fraction of the speed of light. The spacetime in their vicinity is highly nonlinear, dynamical, warped and nontrivial. Numerical relativity runs into problems however, when the separation of the bodies is too large. In theory, this configuration should be tangible, but the issue is due to a matter of practicality. When the separation is increased, evolving the system over many orbits until it approaches the merging phase takes too long. Re- solving over a long orbiting timescale is therefore too computationally taxing.

Numerical relativity also runs into problems when the mass ratio between the two bodies is large. The computational techniques are suited to comparable mass bodies, but when the masses differ by orders or magnitude, we must re- solve over the two disparate length scales, the large and the small, in the system.

Again, this issue is a matter of practice, as this is also computationally taxing.

As the mass ratio is increased, this becomes computationally prohibitive.

So what can be done when the separation of the bodies is large? In this configuration, the bodies are far apart and the orbital velocity is low. The orbit is yet to reach the rapid, highly relativistic merging phase. We say that this regime is in the weak field, and as such, we can make approximations. The weakness of the gravitational field allows us to decompose the metric into the flat Minkowski spacetime ηαβ with a small perturbation pαβon top of it,

gαβ= ηαβ+ pαβ (1.2)

where |pαβ| << 1. This is the regime of validity for post-Minkowskian and post- Newtonian theory. We are essentially asking, how much are we deviating from flat, Minkowski spacetime, and/or from slow, Newtonian orbits? In practice, coalescing binary black holes can be described by an expansion scheme that utilizes both regimes; we expand in powers of G (emphasizing the weak field) as we expand in powers of c−2 (emphasizing slow orbital velocities). This is because in bound gravitating systems, the gravitational potential energy must be equal to the kinetic energy of the system,

G ∼ (v/c)2≈ 1/c2 (1.3)

(7)

which necessitates expanding in both parameters, in a mutually consistent man- ner. These perturbative approximations, although limited in the range of orbital separation and velocity, are not limited in the mass ratio.

Finally, we arrive at the remaining regime of the two-body problem. LIGO is not sensitive to the frequency of gravitational waves produced from such binaries, instead we turn to the future space based detector eLISA. The mass ratio is very large, and the separation between the bodies may be small, so we are in the strong field. Given that one body is far more massive than the other, we take the gravitational field of the large body to be the known metric of a Schwarzschild or Kerr black hole solution to the EFE, and we assume the much smaller body to be a perturbation on top of gαβ0 , the known background spacetime, such that

gαβ= g0αβ+ pαβ (1.4)

The small expansion parameter chosen is the mass ratio, or the small mass µ.

This is the domain of black hole perturbation theory. This regime is suited to perturbations in the strong field of the central body, but also not limited in the orbital separation. The techniques employed can be extended to large distances. Again, as a matter of practicality, we may not want to go to very large separations, as the techniques employed for dealing with the strong field specifically, can be substituted for post-Newtonian methods which are more appropriate in that regime. Although black hole perturbation theory can be pushed far towards large separations (within reason for computational practi- cality), it cannot be pushed towards a comparable mass ratio. The starting assumption of a small body acting as a small perturbation on a highly curved background solution would then no longer be accurate.

Different techniques are therefore employed in their respective domains of validity, but the boundaries between them are not sharply defined. Indeed, the different regimes can be reconciled with each other at the boundaries, and work is actively being done to push one regime closer to the other, and to use results from one domain to inform progress on the other. It should be noted that a further regime, known as effective-one-body (EOB) sets out to encompass arbitrary separations and mass ratios, by combining results from the different domains. This combines analytical and numerical results, and fitting with phenomenological models to produce waveform templates for the inspiral, merger, and finally ringdown phase of coalescing binary compact bodies [4].

1.3 Tidal distortion of Newtonian fluid body

Black holes, singularities in the fabric of spacetime geometry, are remarkable artifacts of Einstein’s celebrated theory of General Relativity. Although black holes do not exist in Newtonian theory, can we draw any parallels with our Newtonian intuition, in our study of tidal deformability of a black hole in GR?

(8)

In Newtonian graivtational theory, we must consider fluid objects. A spher- ical fluid object will be deformed into an ellipsoid by the tidal force exerted from an external, perturbing companion object [5]. If the central fluid body is non-rotating, and the perturbing body is stationary, the induced bulge (location of greatest distortion) points towards the perturbing body that caused it. This is intuitively sensible.

If however, the perturbing body is an orbiting companion, and we assume the central fluid object to have some viscosity ν, the tidal bulge ”lags” the orbiting body by a time interval

τ ∝ Rν

M (1.5)

where R and M are the fluid object’s average radius and mass, respectively [6].

We note that if the fluid body is instead rotating, the bulge will lag behind the orbiting companion, if the spin frequency is lower than the orbital frequency.

If the spin frequency is higher than the orbital frequency, the bulge will instead

”lead” the orbiting body.

For the purpose of comparison, we will restrict our attention to the case in which the fluid body is non-rotating, and hence the distortion lags behind the orbit. How will the event horizon of a non-rotating, Schwarzschild black hole respond to an orbiting companion?

1.4 Tidal distortion of compact objects

The effect of tides on compact objects is an interesting topic of astrophysical study. This is of particular interest in binary systems, in which the bodies may distort one another as the orbit about each other. Therefore, the goal of detect- ing the tidal imprint on gravitational waves from these extreme gravity objects has been pushing research to uncover the precise dynamics of these systems. In the case that one or both of the members in the binary is a neutron star, the prospect of learning about this objects equation of state is particularly enticing (see e.g. [7]). Among other research, some investigations have been concerned with the tidal distortion of black holes, such as the case of the non-rotating (Schwarzschild) (e.g. [8], [9]) or rotating (Kerr) (e.g. [5], [10], [11]) black holes of general relativity. A topic that is of critical importance in such studies is how one chooses to quantify the distortion, given the choice of coordinates describing the geometry of the system. In the case of non-rotating black holes, a gauge in- dependent measurse, the so-called Love numbers, are vanishing quantities [12].

Therefore, a way of quantifying the distrotion, that does not vanish, or rely on the choice of gauge, has been desired. A resolution to this problem has been presented in [6].

In this project, we investigate how a tidally deformed black hole can be de- scribed by the intrinsic and extrinsic geometry of its event horizon by closely following the approach in [6], in particular. In our investigation of the geometry of a deformed horizon, we undertake a mathematical study of the Gauss-Codazzi

(9)

theory of a two-dimensional null hypersurface embedded in the external, four- dimensional spacetime manifold. This is an extension of the well-known de- scription of spacelike or timelike hypersurfaces (e.g [13]). Again, we conduct our study with minimal assumptions as we develop the necessary quantities in differential geometry.

We focus on the tidal distortion of the Schwarzschild spacetime solution to the Einstein field equations of general relativity. The assumption we make is that the tidal distortion is small enough to be described to first (linear) order in perturbation theory. Aside from this requirement, the formalism we study presents no greater restrictions. Following this, we investigate the tidal defor- mation in the regime of black hole perturbation theory. We will discuss the use of master-functions in a gauge-invariant framework [14] that describe the perturbation in the even and odd-parity sectors, in which the metric perturba- tion is decomposed into even and odd-parity spherical harmonics, respectively.

By closely following the approach in [5], we will investigate how to express a measure of tidal deformability in a way that is amenable to visualization, so that we may plot our results in Euclidean space, and clearly discern the tidal effect of a small orbiting body on the Schwarzschild black hole.

Central questions that we aim to answer are therefore: What can we learn from a thorough study of the differential geometry describing the event horizon?

Can we derive meaningful measures of tidal deformability? How does the event horizon respond to a small perturbation? How does it compare with our New- tonian intuition? How can we use black hole perturbation theory to understand the tidal distortion? How can we visualize the tidal influence of an orbiting body, on the Schwarzschild black hole? How does the tidal distortion differ for orbits at different separations?

1.5 Outline of project

This project can be seen as divided into these main categories: geometry, per- turbation theory, and numerical results. Here, we provide a brief outline of the project.

In Sec. 1, we provided an introduction to tidally deformed black holes, by noting that gravitational wave astronomy gives context and motivation for their study, and that we may study them within the two-body problem in general rel- ativity. We noted some of the aspects of tidal distortion of compact objects in the literature, and provided and overview of what our physical intuition from Newtonian theory provides.

In Sec. 2, having set the stage for our investigation, we begin our study of the geometry of null hypersurfaces embedded in spacetime. We begin by carefully detailing the formalism and then use it to derive the Gauss-Weingarten equa- tions, which also lead us to notions of intrinsic and extrinsic quantities, along

(10)

with notions of curvature. These relations are used to derive the Gauss-Codazzi equations, which are then used to obtain components of the Ricci tensor, which ultimately lead us to several measures of tidal deformability.

In Sec. 3, we introduce the Schwarzschild metric and a small perturbation thereof, which sets the focus for later sections. We set up coordinates that en- able the study of a distorted horizon, leading us to horizon equations for the perturbations. We also choose to focus on particular extrinsic and intrinsic ge- ometrical quantities identified earlier, and study them in the context of linear perturbation theory.

In Sec. 4, we introduce techniques from black hole perturbation theory. We provide an overview of master-functions that are well-known in literature and can be related to our geometrical quantities on the horizon. From a near-horizon analysis, we obtain such a relation for the Ricci scalar curvature. We also pro- vide an overview of the Teukolsky equation for Kerr black holes, which we can apply to the scalar curvature in our Schwarzschild study.

In Sec. 5, we study the tidal influence on the Schwarzschild black hole, by a small body in a circular, equatorial orbit. We discuss how the orbital energy has been reconciled with tidal deformability, and provide details of how to embed the horizon in Euclidean space. We also provide a calculation of the dephasing angle between the tidal bulge and the orbiting body, for two different orbital separations.

In Sec. 6, we describe the Teukolsky code which produces the numerical data which we use. We provide plots of the mode convergence of the perturbed scalar curvature, along with snapshots of the tidal bulge at various angular dis- placements of the orbiting body, for both orbital separations that we consider, allowing us to compare the results.

In Sec. 7, we provide a summary of the project, and outline numerous ways to extend this work, and numerous new avenues for investigation. We then pro- vide some concluding remarks.

In App. A, we investigate the effect of reparameterizing the horizon’s null generators. In App. B, we discuss the scalar, vector, and tensor spherical har- monics. In App. C, we provide and a detailed calculation of the induced metric.

In App. D, we provide a detailed calculation of an extrinsic horizon quantity.

Throughout this work, we use units in which G = c = 1.

(11)

2 Hypersurface geometry

Given a spacetime manifold, a hypersurface is a submanifold. A hypersurface is characterized as being either spacelike, timelike or null, with the latter sub- ject to subtleties and requiring special care in its analysis. This is precisely the case that we are interested in when studying the event horizon of a black hole.

It is therefore worth taking some time to develop the appropriate formalism [15].

We may for instance consider a codimension 1 manifold, that is, a 3-dimensional hypersurface Σ embedded in 4-dimensional spacetime manifold M. This would be the case of a spacelike or timelike hypersurface, which may be described in parametric form as

xα= xα(ya) (2.1)

for spacetime coordinates xα and coordinates on the hypersurface ya, and for lower-case Latin indices a = 1, 2, 3, or by a function Φ placing a restriction on the spacetime coordinates for some constant C,

Φ(xα) = C. (2.2)

Now, to study the geometry of a hypersurface, we need the notion of a unit normal vector to the surface. This (normalized) normal, nα is defined to be

nα= ± ∂αΦ

p|gµνµΦ∂νΦ| (2.3)

where ± denotes a timelike (+) or spacelike (−) hypersurface. This notion is not applicable for the null case, as the denominator is zero. Instead, we define

kα= −∂αΦ (2.4)

to be the normal. Given that the event horizon of a black hole is described by a null hypersurface, this is the case we will pursue further.

2.1 Constructing a null hypersurface

The first steps to be taken, with the goal of formulating the geometry of the event horizon, is to discuss the congruence of null geodesics that generate the null hypersurface Σn. A congruence of geodesics is essentially a family of curves that do not intersect. We closely follow the description and notation as pre- sented in [6] to set the formalism to follow.

In the null case, we introduce the parametric equations xα as

xα= xα(λ, αA), αA= (α, β) (2.5) where λ is a scalar parameter along each generator, and αA are for A = 2, 3 labels on the generators that remain constant on each generator, which will

(12)

be more suitable than the form Eq. (2.1). With this, the null vector field introduced previously can be written as

kα= ∂xα

∂λ



αA

(2.6) is both normal and lies tangent to the congruence. The spacelike displacement vectors are similarly defined as

eαA= ∂xα

∂αA



λ

(2.7) These basis vectors satisfy the orthogonality conditions

kαkα= 0 , kαeαA= 0 (2.8)

Since kαis defined to be the normal to the Σn, the first condition is a statement that it is orthogonal to itself, hence also tangent to Σn. This shows the subtleties involved in the null case. Now, as we are dealing with a null hypersurface, the intrinsic geometry (i.e geometry on the surface) is necessarily described by a degenerate two-dimensional metric, which gives the relation between two spacelike displacement vectors as

γAB = gαβeαAeβB (2.9)

where we have contracted over spacetime indices and γAB is taken to be this degenerate 2-metric on Σn and gαβ is the spacetime metric. We let ∇A repre- sent the covariant derivative compatible with the hypersurface metric γAB with connection coefficient given by ΓCAB. From the above definitions, the vectors kα and eαAsatisfy the Lie-transport equations

LeAkα= LkeαA= 0 (2.10a) LeBeαA= LeAeαB= 0 (2.10b) which clearly imply that

eβAβkα= kββeαA (2.11a) eβBβeαA= eβAβeαB (2.11b) where ∇α, the covariant derivative in spacetime, is taken with respect to a connection that is metric compatible with gαβ.

In order to have a full four-dimensional spacetime basis, we complete this set of basis vectors by introducing an auxiliary null vector Nαsuch that

NαNα= 0 , Nαkα= −1 , NαeαA= 0 (2.12) where the second condition is a typical normalization choice, and the third follows for the same reason as for kα, namely that Nαlies tangent to eαA.

Now that we have constructed a complete vector basis on Σn, a natural ques- tion to ask now is how does one differentiate a vector Aαon the hypersurface.

(13)

2.2 Derivation of Gauss-Weingarten equations

In order to arrive at the Gauss-Codazzi equations that we seek to study the ge- ometry of the black hole’s event horizon, we first require the Gauss-Weingarten equations. These serve to express the rate of change of the unit normal vector to a given surface using the derivatives of the position vector on the surface.

Given that the null vector Nαis auxiliary, this leaves us seeking such equations for kββkα, eβAβkαand eβBβeαA.

The first Gauss-Weingarten equation is easily obtained, as it is simply given by the general form of the geodesic equation [15], namely

kββkα= κkα (2.13)

for some scalar κ. An appropriate choice is one that preserves the notion that kα is null on the hypersurface, so we seek no terms proportional to Nαon the RHS. This leads us to find

κ = −Nαkββkα (2.14)

to be an appropriate choice. The remaining relations require more work to elu- cidate.

Consider first the case of spacelike or timelike hypersurfaces. The quantity defined as

DbAa := eαaeβbβAα (2.15) for a tangent vector Aa where lower-case Latin indices denote intrinsic coordi- nates, is the intrinsic covariant derivate. It describes the tangential components of the vector eβbβAα on Σ. Given that we are working with Σn, we seek to find an analogue for null hypersurfaces.

To do this, we require the normal component of the vector defined above.

We start by re-writing the vector as

eβbβAα= gαµeβbβAµ (2.16) where we can now decompose the metric into its normal and tangential parts, using the metric on Σn, so the spacetime metric for the null case is [15]

gαβ= −kαNβ− Nαkβ+ γABeαAeβB (2.17) We proceed as follows

eβBβAα= gαµeβBβAµ

= (−kαNµ− Nαkµ+ γAMeαAeM µ)eβBβAµ

= −(NµeβBβAµ)kα− (kµeβBβAµ)Nα+ γAM(eM µeβBβAµ)eαA (2.18)

(14)

From the tangent vector fields in our basis, we obtain the remaining Gauss- Weingarten equations in the following, by setting Aα = kα and Aα = eαA, respectively.

eβAβkα= ωAkα+ BABeαB (2.19a) eβBβeαA= BABNα+ KABkα+ ΓCABeαC (2.19b) where, as we have noted previously, the Lie transport equations Eq. (2.11a) and Eq. (2.11b) hold, and we have defined

ωA= −NαeβAβkα (2.20a) BAB= eαAeβBβkα (2.20b) KAB= −NαeβBβeαA (2.20c) ΓCAB= eeβBβeαA (2.20d) where BAB= BBA, KAB = KBAand ΓCAB = ΓCBAare symmetric in AB. To show that BAB is symmetric, for example, we proceed as follows

BAB= eαAeβBβkα

= eβBβ(kαeαA) − eβBkαβeαA

= −kαeβAβeαB

= −eβAβ(kαeαB) + eβAeαBβkα

= eαBeβAβkα

= BBA

(2.21)

where the second line follows from the product rule for the covariant deriva- tive, the third line follows from orthogonality and the Lie transport equation Eq. (2.11b), Eq. (2.8) and the remainder of the calculation proceeds by the same set of arguments. Symmetry of KAB follows from a very similar set of arguments.

We can see from Eq. (2.20) that the geometric quantities γAB, BAB and ΓCAB are related to the intrinsic geometry of Σn, whereas κ, ωA, and KAB in contrast, are related to the extrinsic geometry of Σn, i.e how this hypersurface is embedded in the external four-dimensional spacetime.

2.3 Intrinsic and extrinsic quantities

We give some further consideration to the geometric quantities that we have encountered. Now, we further note that

BAB= eαAeβBβkα

= 1

2eαAeβBLkgαβ (2.22)

(15)

represents the intrinsic curvature of the null hypersurface, and we have used a helpful result that relates the covariant derivative of kαto the Lie derivative of the metric along kα[3], [15]. Since kαis null and tangent to the hypersurface, Lkgαβ is a tangential derivative.

We may then wonder what the quantity KAB represents. We write KAB= 1

2eαAeβBLNgαβ

= 1

2eαAeβB(∇βNα+ ∇αNβ)

= −NαeβBβeαA

(2.23)

and define this quantity as the transverse curvature of the null hypersurface, using similar techniques as in the aforementioned calculation, and note that it represents the transverse derivative of the metric.

To conclude this section, we note that these quantities, intrinsic or extrin- sic, all refer to a specific parameterization choice (λ, αA) of the horizon’s null generators. A demonstration of the effect of reparamterization is given in App.

A.

2.4 Derivation of Gauss-Codazzi equations

We mentioned that for spacelike or timelike hypersurfaces, the notion of an intrinsic derivative can be defined. From such a derivative, we may further define the Riemann tensor for the hypersurface as

RcdabAd= DbDaAc− DaDbAc (2.24) in which this curvature tensor is entirely intrinsic. Returning our attention to the null case at hand, the situation is not as straightforward. We know that the Riemann tensor Rαβγδ describes the curvature of a 4D spacetime manifold.

Given that γAB acts as a metric on Σn, we now wish to consider how the 2D Riemann curvature tensor on Σn can be related to the 4D curvature tensor.

This will ultimately serve to bring us closer to understanding how the geometry of a black hole’s event horizon is described, and how it is then altered by an external influence.

To arrive at the Gauss-Codazzi equations, we follow Sec. 3.5 of [15], and begin by taking the covariant derivative along eγCof the Gauss-Weingarten equa- tions. Consider the first such equation given by Eq. (2.13). We have

eγCγ(kββkα) = eγCγ(κkα) (2.25) and we consider each side of the equality separately. Consider first the LHS,

(16)

which gives

eγCγ(kββkα) = eγCγkβ(∇βkα) + ∇βkα(eγCγkβ)

= eγCkβγβkα+ ∇βkαCkβ+ BCDeβD)

= eγCkβγβkα+ kββkαωC+ BCDeβDβkα

= eγCkβγβkα+ kββkαωC+ BCDeβDβkα

= eγCkβγβkα+ κkαωC+ BCDDkα+ BDEeαE)

(2.26)

where we have used the product rule to obtain two terms, then used the Gauss- Weingarten equation Eq. (2.19a) to rewrite the second term. We then expanded this second term, and used the Gauss-Weingarten equations Eq. (2.13) and Eq.

(2.19a) to rewrite the resulting first and second terms, respectively.

We follow similar steps to expand out the RHS and write the appropriate combinations of terms as Gauss-Weingarten relations,

eγCγ(κkα) = kαeγCγκ + κeγCγkα

= kαeγCγκ + κ(ωCkα+ BCDeαD) (2.27) With these expansions at hand, we set LHS = RHS and solve for the term eγCkβγβkα. The reason we are interested in this term will become apparent shortly. We find

eγCkβγβkα= −κkαωC− BCDDkα+ BDEeαE) + kαeγCγκ + κ(ωCkα+ BCDeαD)

= −BCDDkα+ BDEeEα − κeαD) + kαeγCγκ

(2.28) where on the first line, the first terms cancels out, and we are able to re-factorize.

Recall now the form of the Reimann tensor. If we subtract an analogous equation for kβeγCβγkα, we get precisely −Rµβγα kµkβeγC. We find that

RαµβγkµkβeγC= kβeγCβγkα− eγCkβγβkα

= BCDDkα+ BDEeEα − κeαD) − kαeγCγκ + kαkββωC+ eαDkββBCD

(2.29)

What is left to do now is to project along the various basis vectors. Projecting along Nαgives

RαµβγkµkβeγCNα= −BCDωD+ eγCγκ − kββωC

= −BCDωD+ ∂λωC− ∂Cκ (2.30) where we recall that κ is a scalar, and we have used Eq. (2.6) to write the last covariant derivative as ∂λ. If we instead project along e, the resulting

(17)

equation is

RαµβγkµkβeγCe= BCDBDEeαEe− κBCDeαDe+ eαDekββBCD

= BCDBCDBDA− κBCA+ ∂λBCA

(2.31) Following similar steps with the remaining Gauss-Weingarten equations, and considering the various projections along our basis vectors, we arrive at a set of equations which define the Gauss-Codazzi relations, listed below.

kµNνkλeαARµνλα= ∂λωA− ∂Aκ + BABωB (2.32a) kµeαAkνeβBRµανβ= −∂λBAB− κBAB+ BACBCB (2.32b) kµeαANνeβBRµανβ= −∂λKAB− κKAB+ ∇AωB+ ωAωB+ KACBCB (2.32c) kµNνeαAeβBRµναβ= ∇AωB− ∇BωA− BACKCB+ BBCKCA (2.32d) kµeαAeβBeγCRµαβγ= ∇CBAB− ∇BBAC− ωCBAB+ ωBBAC (2.32e) NµeαAeβBeγCRµαβγ= ∇CKAB− ∇BKAC+ ωCKAB− ωBKAC (2.32f) eαAeβBeγCeδDRαβγδ= ¯RABCD+ BACKBD− BADKBC+ KACBBD− KADBBC

(2.32g) These relations are the various projections of the spacetime Riemann tensor that we are able to express in terms of the intrinsic and extrinsic curvature quantities on the null hypersurface. The relation Eq. (2.32g) directly relates the spacetime Riemann tensor to the Riemann tensor ¯RABCDon the null hypersurface Σn.

2.5 Components of Ricci tensor

With this set of Gauss-Codazzi equations, we are now equipped go on to find the components of the spacetime Ricci tensor. We begin by writing

Rµν = gαβRαµβν

= (−kαNβ− Nαkβ+ γABeαAeβB)Rαµβν

= −RαµβνkαNβ− RαµβνNαkβ+ γABRαµβνeαAeβB

(2.33)

where at the first equality, we have written the Ricci tensor as a the spacetime metric Eq. (2.17) contracted with the Riemann tensor; at the second equality we inserted the form of this metric; and at the last equality we expanded out to write as Riemann tensors projected along two basis vectors. Projecting along a further two basis vectors allows us to find the components of the Ricci scalar.

(18)

The first component we find is

Rµνkµkν = −RαµβνkαkµNβkν− RαµβνNαkµkνkβ+ γABRαµβνeαAkµeβBkν

= γABRαµβνeαAkµeβBkν

= γABRµανβkµeαAkνeβB

= γAB(−∂λBAB− κBAB+ BACBCB)

= −∂λΘ + κΘ + BABBAB

(2.34) where we have used the symmetry in the indices µ, α and α, ν to cancel the first and second terms, respectively; rearranging the indices on the remaining term allows us to relate it directly to Eq. (2.32b); contracting with the two-metric then gives us our result where we defined the scalar trace

Θ := γABBAB (2.35)

which denotes the expansion of the null geodesic congruence. Recall that a con- gruence of null geodesics generates the event horizon, so this quantity is a useful measure of how the event horizon expands or contracts. We will have more to say about this in a later section.

The second component of the Ricci tensor is found from similar steps, and we find

RµαkµeαA= −RµανβkµeαAkνNβ− RµανβkµeαANνkβ+ γN BRµανβkµeαAeνNeβB

= −RµνλαkµNνkλeαA+ γBCRµαβγkµeαAeβBeγC

= ∂λωA− ∂Aκ + BABωB+ γBC(∇CBAB− ∇BBAC− ωCBAB+ ωBBAC)

= ∂λωA− ∂Aκ + BABωB+ ∇BBAB− ∇AΘ − ωBBAB+ ωBΘ

= ∂λωA− ∂Aκ − ∂AΘ + ωBΘ + ∇BBAB

(2.36) The third component requires slightly more work to elucidate. We begin as before, and write this component of the Ricci tensor as the sum of components of the Riemann tensor,

RαβeαAeβB= −RµανβkµeαANνeβB− RµανβNµeαAkνeβB+ γCDRµανβeµCeαAeνDeβB (2.37) where on the first line, we have used the fact that the first term is given by Eq.

(2.32c) and the third term is composed of the induced metric and Eq. (2.32g).

We now use the first Bianchi identity,

Rµ[ανβ]= 0 (2.38)

(19)

to rearrange and rewrite the second term as

RµανβNµkνeαAeβB= −RµνβαNµkνeαAeβB− RµβανNµkνeαAeβB

= −RνµβαkµNνeαAeβB− RνβαµkµNνeαAeβB

= −RµναβkµNνeαAeβB− RµανβkµNνeαAeβB

(2.39)

where we have rearranged indices on the Riemann tensors, and thus expressed the second term as a sum of two known Gauss-Codazzi relations, namely Eq.

(2.32d) and Eq. (2.32c). With this, we may now continue calculating the final component of the Ricci tensor

RαβeαAeβB = −2RµανβkµeαANνeβB+ RµναβkµNνeαAeβB+ γCDRµανβeµCeαAeνDeβB

= 2(∂λKAB+ κKAB− ∇AωB− ωAωB− KACBCB) + ∇AωB− ∇BωA

− BACKCB+ BBCKCA+1 2

Rγ¯ ACACγBD− γADγBC) + γAC(BACKBD− BADKBC+ KACBBD− KADBBC)

= 2(∂λ+ κ + 1

2)KAB− (∇AωB+ ∇BωA) − 2ωAωB

− 2(KACBCB+ KBCBCA) + KBAB+1 2

Rγ¯ AB

(2.40) where at the second equality, we used a geometric property of the Riemann tensor in 2D [6], namely that it can be written as

ABCD=1 2

R(γ¯ ACγBD− γADγBC) (2.41) where ¯R is the degenerate two-dimensioanl Ricci scalar on Σn. In analogy to Eq. (2.35), we have also defined a trace term for KAB for convenience, as

K := γABKAB (2.42)

which we used at the third equality. Although, the quantity K presents no similar insight as the analogous scalar Θ.

To summarize the results of these calculations, we have computed the com- ponents of the Ricci tensor. They involve the intrinsic and extrinsic quantities we have discussed earlier, and they are given as follows

kµkνRµν = −(∂λ− κ)Θ + BABBAB (2.43a)

kµeαARµα= ∂λωA− ∂Aκ − (∂A− ωA)Θ + ∇BBAB (2.43b) eαAeβBRαβ= 2(∂λ+ κ + 1

2)KAB− (∇AωB+ ∇BωA) − 2ωAωB

− 2(KACBCB+ KBCBCA) + KBAB+1 2

Rγ¯ AB

(2.43c)

(20)

where in Eq. (2.43a), we have corrected the sign of the final term to have a plus, as compared with the result stated in [6]. Equation Eq. (2.43a) will be particularly useful in revealing physical insight on the behavior of the event horizon.

2.6 Evolution of expansion and shear

We found previously that the geometric quantity BAB is a characteristic of the intrinsic geometry of our null hypersurface Σn. Furthermore we found that it represents the intrinsic curvature of Σn. We now find further interpretation of this quantity, by instead examining its components. What we may do is to split this tensor into components and write

BAB= 1

2ΘγAB+ σAB (2.44)

where the first term is Eq. (2.35) as before and denotes the expansion of the null geodesic congruence, and σAB is the remaining term. As such, it is purely tracefree and instead denotes the shearing of the null congruence. To see how these quantities change, we desire an evolution equation for the rate of expansion and shear. For the expansion Θ, this is precisely given by Raychaudhuri’s equation [15], which we arrive at by making use of the results recently developed.

Rearranging Eq. (2.43a) and substituting in the decomposition Eq. (2.44), the rate of expansion is

λΘ = κΘ − BABBAB− Rαβkαkβ

= κΘ − (1

2ΘγAB+ σAB)(1

2ΘγAB+ σAB) − Rαβkαkβ

= κΘ −1

2− 2ΘγABσAB− σABσAB− Rαβkαkβ

(2.45)

where at the final equality, we have made use of γABγAB= δAA= 2. We continue the calculation by rearranging Eq. (2.44) and substituting it in for σAB, which enables us to perform a useful cancellation of the term we are taking the trace of at the second equality. This is again due to σAB being purely tracefree.

Continuing the calculation,

λΘ = κΘ −1

2− 2ΘγAB(BAB−1

2ΘγAB) − σABσAB− Rαβkαkβ

= κΘ −1

2− 2ΘγAB(Θ − Θ) − σABσAB− Rαβkαkβ

= κΘ −1

2− σABσAB− Rαβkαkβ

(2.46)

We have now simplified the evolution equation and it is left in terms of a com- ponent of the Ricci tensor. This is a useful relation, however we can do a bit more to obtain a deeper physical interpretation. Therefore, we are now at the

(21)

stage where we wish to introduce the Einstein field equations Eq. (1.1), Gαβ:= Rαβ−1

2Rgαβ= 8πTαβ (2.47)

where we write the Einstein tensor Gαβin terms of quantities we have developed.

Using the Einstein field equations, we may re-express the component of the Ricci tensor in question as

Rαβkαkβ= (8πTαβ+1

2Rgαβ)kαkβ

= 8πTαβkαkβ+1 2R kβkβ

| {z }

=0

(2.48)

where the final term vanishes by virtue of orthogonality Eq. (2.8), so we may now express Raychaudhuri’s equation as

λΘ = κΘ −1

2− σABσAB− 8πTαβkαkβ (2.49) where the term introduced involving the energy-momentum denotes matter moving across Σn. This is our evolution equation for the expansion scalar Θ, which describes the expansion of a null geodesic congruence generating a null hypersurface, which in turn describes the black hole event horizon.

Now, if instead we start from Eq. (2.32b), we obtain a different evolution equation. Rearranging yields,

λBAB = −κBAB+ BACBCB− RµανβkµeαAkνeβB (2.50) which can be written in terms the tracefree component as

λσAB= (κ − Θ)BAB+ BACBCB− RµανβkµeαAkνeβB (2.51) so we arrive at an evolution equation, analogously to that for Θ, for the shear tensor σBA

λσAB= (κ − Θ)σAB− CµανβkµekνeβB (2.52) where Cµανβis the well known Weyl curvature tensor and is the tracefree con- tribution of the Riemann tensor, whereas the trace contribution is given by the Ricci tensor.

The effect of the shear tensor in general, is slightly more complicated than that of the expansion scalar [15]. The shear tensor can be thought of as be- ing made up of components, or shear parameters, σ+ and σ× which dictate the geometry of the shearing effect, which is area preserving. If σ× = 0, the geodesic congruence, if initially thought of as circular, will become an ellipse with its major axis along an angle of φ = 0 for instance. If instead σ+ = 0, then the initially circular configuration will form an ellipse with major axis in the direction φ = π/4 in comparison. The shearing effects of σ+ and σ× thus

(22)

amount to deformation of the initial configuration along φ = 0 and φ = π/2 and along φ = π/4 and φ = 3π/4, respectively, in a ”plus” and ”cross” pattern, as the labeling of the shear parameters suggests. This is similar to the effect of the + and × polarizations of gravitational waves on a ring of test particles [3].

Looking ahead, in Sec. 4.4 we investigate how these measures of tidal de- formability can be developed further using black hole perturbation theory, and how they apply to the case of a small body external to the black hole, which induces the tidal distortion of the black hole.

References

Related documents

46 Konkreta exempel skulle kunna vara främjandeinsatser för affärsänglar/affärsängelnätverk, skapa arenor där aktörer från utbuds- och efterfrågesidan kan mötas eller

Generally, a transition from primary raw materials to recycled materials, along with a change to renewable energy, are the most important actions to reduce greenhouse gas emissions

Both Brazil and Sweden have made bilateral cooperation in areas of technology and innovation a top priority. It has been formalized in a series of agreements and made explicit

För att uppskatta den totala effekten av reformerna måste dock hänsyn tas till såväl samt- liga priseffekter som sammansättningseffekter, till följd av ökad försäljningsandel

The increasing availability of data and attention to services has increased the understanding of the contribution of services to innovation and productivity in

Generella styrmedel kan ha varit mindre verksamma än man har trott De generella styrmedlen, till skillnad från de specifika styrmedlen, har kommit att användas i större

Parallellmarknader innebär dock inte en drivkraft för en grön omställning Ökad andel direktförsäljning räddar många lokala producenter och kan tyckas utgöra en drivkraft

Närmare 90 procent av de statliga medlen (intäkter och utgifter) för näringslivets klimatomställning går till generella styrmedel, det vill säga styrmedel som påverkar