close encounters of black holes - arxiv · 2018. 10. 26. · of black hole collision: •...

31
arXiv:gr-qc/0302111v1 27 Feb 2003 Close encounters of black holes Domenico Giulini Department of Physics University of Freiburg Hermann-Herder-Strasse 3 D-79104 Freiburg Germany Abstract This is an introduction into the problem of how to set up black hole initial- data for the matter-free field equations of General Relativity. The approach is semi-pedagogical and addresses a more general audience of astrophysicists and students with no specialized training in General Relativity beyond that of an introductory lecture. 1 1 Introduction and motivation In my lecture I will try to explain how scattering and merging processes between black holes can be described analytically in general relativity(GR). This is a vast subject and I will focus attention to the basic issues, rather than trying to explain analytical details of approximation schemes etc. I will also not discuss numerical aspects, which are beyond my competence, and which would anyway require a separate lecture. I will address the following main topics: 1) a first step beyond Newtonian gravity, 2) constrained evolutionary structure of Einstein’s equations, 3) the 3+1- split and the Cauchy initial-value problem, 4) black hole data, 5) problems and recent developments, 1 This is the written version of a lecture delivered at the school “The Galactic Black Hole”, held between August 26.-31. in 2001 at the Physics Center of the German Physical Society in Bad Honnef (Germany). It is published in: Heino Falcke and Friedrich W Hehl (editors) The Galactic Black Hole (Lectures on General Relativity and Astrophysics), IOP Publishing (Bristol) 2003. 1

Upload: others

Post on 25-Jan-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

  • arX

    iv:g

    r-qc

    /030

    2111

    v1 2

    7 F

    eb 2

    003

    Close encounters of black holes

    Domenico Giulini

    Department of Physics

    University of Freiburg

    Hermann-Herder-Strasse 3

    D-79104 Freiburg

    Germany

    Abstract

    This is an introduction into the problem of how to set up blackhole initial-data for the matter-free field equations of General Relativity. The approach issemi-pedagogical and addresses a more general audience of astrophysicistsand students with no specialized training in General Relativity beyond thatof an introductory lecture.1

    1 Introduction and motivation

    In my lecture I will try to explain how scattering and mergingprocesses betweenblack holes can be described analytically in general relativity(GR). This is a vastsubject and I will focus attention to the basic issues, rather than trying to explainanalytical details of approximation schemes etc. I will also not discuss numericalaspects, which are beyond my competence, and which would anyway require aseparate lecture. I will address the following main topics:

    1) a first step beyond Newtonian gravity,

    2) constrained evolutionary structure of Einstein’s equations,

    3) the 3+1- split and the Cauchy initial-value problem,

    4) black hole data,

    5) problems and recent developments,1This is the written version of a lecture delivered at the school “The Galactic Black Hole”, held

    between August 26.-31. in 2001 at the Physics Center of the German Physical Society in Bad Honnef(Germany). It is published in: Heino Falcke and Friedrich W Hehl (editors)The Galactic Black Hole(Lectures on General Relativity and Astrophysics), IOP Publishing (Bristol) 2003.

    1

    http://arxiv.org/abs/gr-qc/0302111v1

  • with emphasis on the fourth entry. However, I will also spendsome time to explainsome of the specialties of GR, like the absence of a point-particle concept andthe non-trivial linkage between the field equations and the equations of motion formatter. These points should definitely be appreciated before one goes on to discussblack holes, which are solutions to thevacuumEinstein equations representingextendedobjects.

    The following points seem to me the main motivations for studying the problemof black hole collision:

    • Coalescing black holes are regarded as promising sources for the detectionof gravitational waves by earth-based instruments.

    • Close encounters of black holes provide physically relevant situations for theinvestigation of the strong-field regime of general relativity.

    • The dynamics of simple black hole configurations is regardedas ideal testbedfor numerical relativity.

    My conventions are as follows: space-time is a manifoldM with Lorentzian metricg of signature(−,+,+,+). Greek indices range from0 to3, latin indices from1 to3 unless stated otherwise. The covariant derivative is denoted∇µ, ordinary partialderivatives by∂µ or sometimes simply by a lower-case, µ. The relation:= (=:)defines the left (right) hand side. The gravitational constant in GR isκ = 8πG/c2,whereG is Newton’s constant andc the velocity of light. A symbol likeO(ǫn)stands collectively for terms falling off at least as fast asǫn.

    2 A first step beyond Newtonian gravity

    It can hardly be overstressed how useful the concept of amass-pointis in New-tonian mechanics and gravity. It allows topointwiseprobe the gravitational fieldand to reduce the dynamical problem to the mathematical problem of finding so-lutions to a system offinitely manyordinary differential equations. To be sure,just postulating the existence of mass points is not sufficient. To be consistent withknown laws of physics one must eventually understand the point mass as an ideal-ization of a highly localized mass distribution which obeysknown field-theoreticlaws, such that in the situations at hand most of the field degrees of freedom effec-tively decouple from the dynamical laws for those collective degrees of freedomone is interested in, like e.g. the centre-of-mass. In Newtonian gravity this usuallyrequires clever approximation schemes but is not considered to be a problem offundamental nature. Although this is true for the specific linear theory of Newto-nian gravity, this needs not be so for comparably simple generalizations, as willbecome clear below.

    2

  • In GR the situation is markedly different. A concentration of more than oneSchwarzschild mass in a region of radius less than the Schwarzschild radius willlead to a black hole whose behaviour away from the stationarystate usually can-not be well described by finitely many degrees of freedom. It shakes and vibrates,thereby radiating off energy and angular momentum in form ofgravitational radi-ation. Moreover, it is an extended object and cannot be unambiguously ascribedan (absolute or relative) position or individual mass. Hence the problem of mo-tion, and therefore the problem of scattering of black holes, cannot be expected tomerely consist ofcorrectionsto Newtonian scattering problems. Rather, the wholekinematic and dynamical setup will be different where many of the establishedconcepts of Newtonian physics need to be replaced or at leastadapted, very oftenin a somewhat ambiguous way. Among those are mass, distance,and kinetic en-ergy. For example, one may try to solve the following straightforward soundingproblem in GR, whose solution one might think has been given long ago. Givenare two unspinning black holes, momentarily at rest, with equal individual massm,mutual distanceℓ, and no initial gravitational radiation around. What is theamountof energy released via gravitational radiation during the dynamical infall? In suchsituation we can usually make sense of the notions of ‘spin’ (hence unspinning) and‘mass’; but ambiguities generally exist in defining ‘distance’ and, most importantof all, ‘initial gravitational radiation’. Such difficulties persist over and above theubiquitous analytical and/or numerical problems which arecurrently under attackby many research groups.

    To those who are not so familiar with GR and like to see Newtonian analogies,I wish to mention that there is a way toconsistentlymodel some of the non-linearfeatures of Einstein’s equations in a Newtonian context, which shares the propertythat it does not allow for point masses. I will briefly describe this model since itdoes not seem to be widely known.

    First recall the field equation in Newtonian gravity, which allows to determinethe gravitational potentialφ (whose negative gradient, -~∇φ, is the gravitationalfield) from the mass densityρ (the ‘source’ of the gravitational field):

    ∆φ = 4πGρ . (1)

    Now suppose one imposes the following principle for a modification of (1): allenergies, including the self-energy of the gravitational field, act as source for thegravitational field. In order to convert an energy densityε into a mass densityρ,we adopt the relationε = ρc2 from special relativity (the equation we will arriveat can easily be made Lorentz invariant by adding appropriate time-derivatives).The question then is whether one can modify the source term of(1) such thatρ→ρ+ρgrav with ρgrav := εgrav/c2, whereεgrav is the energy density of the gravitational

    3

  • field as predicted by this very same equation(condition of self-consistency). Itturns out that there is indeed a unique such modification, which reads

    ∆φ =4πG

    c2φ

    [

    ρ+c2

    8πG(~∇φ/φ)2

    ]

    . (2)

    It is shown in the Appendix that this equation indeed satisfies the ‘energy principle’as just stated. (For more information and a proof of uniqueness, see [21].) Thegravitational potential is now required to be always positive, tending to the valuec2 at spatial infinity (rather than zero as for (1)). The second term on the right handside of (2) corresponds to the energy density of the gravitational field. Unlike theenergy density following from (1) (which is− 18πG |∇φ|2) it is nowpositivedefinite.This does not contradict attractivity of gravity for the following reason: the rest-energy density of a piece of matter is in this theory not givenby ρc2, but byρφ, thatis, it depends on the value of the gravitational potential atthe location of matter.The same piece of matter located at lower gravitational potential has less energythan at higher potential values. In GR this is called the universal redshift effect.Here, as in GR, the active gravitational-mass also suffers from this redshift, as isimmediate from the first term on the right hand side of (2), whereρ does not enteralone, as in (1), but is multiplied with the gravitational potentialφ. With respect tothese features our modification (2) of Newtonian gravity mimics GR quite well.

    We mention in passing that (2) can be ‘linearized’ by introducing the dimen-sionless fieldψ :=

    √φ/c, in terms of which (2) reads

    ∆ψ =2πG

    c2ρψ . (3)

    The boundary conditions are nowψ(r → ∞) → 1. Hence only those linearcombinations of solutions are again solutions whose coefficients add up to one. Forρ ≥ 0 it also follows that solutions to (3) can never assume negative values, sinceotherwise the functionψ must have a negative minimum (because of the positiveboundary values) and therefore non-negative second derivatives there. But then (3)cannot be satisfied at the minimum, henceψmust be nonnegative everywhere. Thisimplies that solutions of (2) are also non-negative. To be sure, for mathematicalpurposes (3) is easier to use than (2), but note thatφ and notψ is the physicalgravitational potential.

    We now show how these non-linear features render impossiblethe notion ofpoint mass, and even induce a certain black-hole behaviour on their solutions. Letus be interested in static, spherically symmetric solutions to (2) with sourceρ,which is zero forr > R and constant forr < R. We need to distinguish twonotions of mass. One mass just counts the amount of ‘stuff’ located withinr < R.

    4

  • You may call it ‘bare mass’ or ‘baryonic mass’, since for ordinary matter it isproportional to the baryon number. We denote it byMB . It is simply given by

    MB :=

    space

    d3x ρ . (4)

    The other mass is the ‘gravitational mass’, which is measured by the amount offlux of the gravitational field to ‘infinity’, that is, throughthe surface of a spherewhose radius tends to infinity. We call this massMG. It is given by

    MG :=1

    4πGlimr→∞

    S2(r)(−~∇φ · ~n) dσ , (5)

    wherer = |~x|, ~x/r = ~n, S2(r) is the 2-sphere or radiusr anddσ is its surfaceelement.MG should be identified with the total inertial mass of the system, in fullanalogy to the ADM-mass in GR (see eq. (35) below). HenceMGc2 is the totalenergy of the system, with gravitational binding energy also taken into account.The massesMB andMG can dimensionally be turned into radii by writingRB :=GMB/c

    2 andRG := GMG/c2, and further turned into dimensionless quantitiesvia rescaling withR, the radius of our homogeneous star. We writex := RB/Randy := RG/R.

    For each pair of values for the two parametersMB andR there is a uniquehomogeneous-star-solution to (2), whose simple analytical form needs not inter-est us here (see [21]). Using it we can calculateMG, whose dependence on theparameters is best expressed in terms of the dimensionless quantitiesx andy:

    y = f(x) = 2

    (

    1− tanh(√

    3x/2)√

    3x/2

    )

    . (6)

    The functionf maps the interval[0,∞] monotonically to[0, 2]. This implies thefollowing inequality

    MG < 2Rc2/G (7)

    which says that the gravitational mass of the star is boundedby a purely geometricquantity. It corresponds to the statement in GR that the star’s radius must be big-ger than its Schwarzschild radius, which in isotropic coordinates is indeed givenby RS = GMG/2c2. It can be proven [21] that the bound (7) still exists fornon-homogeneous spherically symmetric stars, so that the somewhat unphysicalhomogeneity assumption can be lifted. The physical reason for this inequality isthe ‘redshift’, i.e. the fact, that the same bare mass at lower gravitational poten-tial produces less gravitational mass. Hence adding more and more bare mass intothe same volume pushes the potential closer and closer to zero (recall thatφ is

    5

  • always positive) so that the added mass becomes less and lesseffective in gener-ating gravitational fields. The inequality then expresses the mathematical fact thatthis ‘redshifting’ is sufficiently effective so as to give finite upper bounds to thegravitational mass, even for unbounded amounts of bare mass.

    The energy balance can also be nicely exhibited. Integrating the matter en-ergy densityφρ and the energy density of the gravitational field,c

    4

    8πG (~∇φ/φ)2, we

    obtain

    Ematter = MBc2

    (

    1− 6RB5R

    +O(R2B/R2)

    )

    , (8)

    Efield = MBc2

    (

    3RB5R

    +O(R2B/R2)

    )

    , (9)

    Etotal = MBc2

    (

    1− 3RB5R

    +O(R2B/R2)

    )

    =MGc2 . (10)

    Note that the term−3MBc2RB/5R in (10) is just the Newtonian binding energy.At this point it is instructive to verify the remarks we made above about the posi-tivity of the gravitational energy. Shrinking a mass distribution enhancesthe fieldenergy, but diminishes the matter energy twice as fast, so that the overall energyis also diminished, as it must due to the attractivity of gravity. But here this isachieved with all energies involved being positive, unlikein Newtonian gravity.Note that the total energy,MG, cannot become negative (sinceφ cannot becomenegative, as already shown). Hence one also cannot extract an infinite amount ofenergy by unlimited compression, as it is possible in Newtonian gravity. This isthe analogue in our model theory to the positive mass theoremin GR.

    We conclude by making the point announced above, namely thatthe inequality(7) shows that point objects of finite gravitational mass do not exist in the the-ory based upon (2); mass implies extension! Taken together with the lesson fromspecial relativity, that extended rigid bodies also do not exist (since the speed ofelastic waves is less thanc), we arrive at the conclusion that the dynamical prob-lem of gravitating bodies and their interaction is fundamentally of field-theoretic(rather than point-mechanical) nature. Its proper realization is GR to which wenow turn.

    3 Constrained evolutionary structure of Einstein’s equa-tions

    In GR the basic field is the spacetime metricgµν , which comprises the gravitationaland inertial properties of spacetime. It defines what inertial motion is, namely a

    6

  • geodesicẍλ + Γλµν ẋ

    µẋν = 0 (11)

    with respect to the Levi-Civita connection

    Γλµν :=12g

    λσ(−gµν,σ + gσµ,ν + gνσ,µ) . (12)

    (Since inertial motion is ‘force-free’ by definition, you may rightly ask whetherit is correct to call gravity a ‘force’.) The gravitational field gµν is linked to thematter content of spacetime, represented in form of the energy momentum tensorTµν , by Einstein’s equations

    Gµν := Rµν − 12gµνR = κTµν . (13)

    Due to the gauge-invariance with respect to general differentiable point transforma-tions (i.e. diffeomorphisms) of spacetime one has theidentities(as a consequenceof Noether’s 2nd theorem)

    ∇µGµν ≡ 0 . (14)Being ‘identities’ they hold for anyGµν , independent of any field equation. Withrespect to some coordinate systemxµ = (x0, · · · x3) we can expand (14) in termsof ordinary derivatives. Preferring the coordinatex0, this reads

    ∂0G0ν = −∂kGkν − ΓµµλGλν − ΓνµλGµλ. (15)

    SinceGµν contains no higher derivatives ofgµν than the second, the right-handside of this equation also contains only 2ndx0-derivatives. Hence (15) impliesthat the four componentsG0ν only involve firstx0-derivatives. Now choosex0 astime coordinate. The four(0, ν)-components of (13) then do not involve secondtime derivatives, unlike the space-space components(i, j). Hence the time-timeand time-space components areconstraints, that is, equations that constrain theallowed choices of initial data, rather than evolving them.

    That is not an unfamiliar situation, which similarly occursfor Maxwell’s equa-tions in electromagnetism (EM). Let us recall this analogy.We consider the four-dimensional form of Maxwell’s equations in terms of the vector potentialAµ,whose antisymmetric derivative is the field-tensorFµν := ∂µAν −∂νAµ, compris-ing the electric (Ei = F0i) and magnetic (Bi = −Fjk, ijk cyclic) field. Maxwell’sequations are

    Eν := ∂µFµν − 4π

    cjν = 0 . (16)

    Due to its antisymmetry, the field tensor obviously obeys theidentity

    ∂µ∂νFµν ≡ 0 , (17)

    7

  • which here is the analogue of (14), an identity involving third derivatives in thefield variables. Using (17) in the divergence of (16) yields

    ∂νEν = −4π

    c∂νj

    ν , (18)

    which shows that Maxwell’s equations imply charge conservation as integrabil-ity condition. Let us interpret the rôle of charge conservation in the initial valueproblem. Decomposing (17) into space and time derivatives gives

    ∂0∂νF0ν = −∂k∂νF kν . (19)

    Again the right hand side involves only second time derivatives implying that the0-component of (16) involves no second time derivatives. Hence the time componentof (16) is merely a constraint on the initial data; clearly itis just Gauss’ law~∇ ·~E − 4πρ = 0. Its change under time evolution according to Maxwell’s equationsis

    ∂0E0 = ∂νE

    ν − ∂kEk

    = −4πc∂νj

    ν − ∂kEk , (20)

    where we used the identity (18) in the second step. Suppose now that on the initialsurface of constantx0 we put an electromagnetic field which satisfies the con-straint,E0 = 0, and which we evolve according toEk = 0 (implying ∂kEk = 0on that initial surface). Then (20) shows that charge conservation is the necessaryand sufficient condition for the evolution to preserve the constraint.

    Let us return to GR now, where the overall situation is entirely analogous. Nowwe have four constraints

    E0ν := G0ν − κT 0ν = 0 (21)

    and six evolution equations, which we write

    Eij := Gij − κT ij = 0 . (22)

    The identity (14) now implies

    ∇µEµν = −κ∇µT µν , (23)

    which parallels (18). Here, too, the time derivative of the constraints is easilycalculated:

    ∂0E0ν = ∇µEµν − ∂kEkν − Γ00λEλν − Γν0λE0λ

    = −κ∇µT µν − ∂kEkν − Γ00λEλν − Γν0λE0λ (24)

    8

  • using (23) in the last step. Now consider again the evolutionof initial data froma surface of constantx0. If they initially satisfy the constraints and are evolvedvia Eij = 0 (hence all spatial derivatives ofEµν vanish initially) they continueto satisfy the constraints if and only if the energy momentumtensor of the mattersatisfies

    ∇µT µν = 0 . (25)Hence we see that the ‘covariant conservation’ of energy-momentum, expressed by(25), plays the same rôle in GR as charge conservation playsin EM. This meansthat you cannot just prescribe the motion of matter and then use Einstein’s equa-tions to calculate the gravitational field produced by that source. You have to movethe matter in such a way that it satisfies (25). But note that atthis point there is acrucial mathematical difference to charge conservation inEM: charge conservationis a condition on the sourceonly, it does not involve the electromagnetic field. Thismeans that you knowa priori what to do in order not to violate charge conserva-tion. On the other hand, (25) involves the sourceand the gravitational field. Thelatter enters through the covariant derivatives which involve the metricgµν throughthe connection coefficients (12). Hence here (25) cannot be solved a priori by suit-ably restricting the motion of the source. Rather we have a consistency conditionwhich mutually links the problem of motion for the sources and the problem offield determination. It is this difference which makes the problem of motion inGR exceedingly difficult. (A brief and lucid presentation ofthis problem, drawingattention to its relevance in calculating the generation ofgravitational radiation byself-gravitating systems, was given in [15]. A broader summary, including moderndevelopments is [14]).

    For example, for pressureless dust represented byT µν = ρc2 UµUν , whereρ is the local rest-mass density andUµ is the vector field of four-velocities ofthe continuously dispersed individual dust grains, (25) isequivalent to the twoequations

    ∇µ(ρUµ) = 0 , (26)Uν∇νUµ = 0 . (27)

    The first states the conservation of rest-mass. The second isequivalent to the state-ment that the vector fieldUµ is geodesic, which means that its integral lines (theworldlines of the dust grains) are geodesic curves (11)with respect to the met-ric gµν . Hence we see that in this case the motion of matter is fully determinedby (25), i.e. by Einstein’s equations, which imply (25) as integrability condition.This clearly demonstrates how the problem of motion is inseparably linked withthe problem of field determination, and that these problems can only be solvedsimultaneously. The methods used today use clever approximation schemes. For

    9

  • example, one can make use of the fact that there is a difference of one power inκbetween the field equations and their integrability condition. Hence, in an approxi-mation inκ it is consistent for thenth order approximation of the field equations tohave the integrability conditions (equations of motions) satisfied ton− 1st order.

    Clearly the problem just discussed does not arise for the matter-free Einsteinequations for whichTµν ≡ 0. Now recall that black holes too are described bythe matter free equations. Hence the mathematical problem just described does notoccur in the discussion of their dynamics. Inthis aspect the discussion of blackhole scattering is considerably easier than e.g. that of neutron stars.

    4 The 3+1- split and the Cauchy initial-value problem

    We saw that the ten Einstein equations decompose into two sets of four and sixequations respectively, four constraints which the initial data have to satisfy, andsix equations driving the evolution. As a consequence therewill be four dynam-ically undetermined components among the ten components ofthe gravitationalfield gµν . The task is to parametrise thegµν in such a way that four dynamicallyundetermined functions can be cleanly separated from the other six. One way toachieve this is via the splitting of spacetime into space andtime (see [22] for amore detailed discussion). The four dynamically undetermined quantities will bethe famouslapse(one functionα) andshift (three functionsβi). The dynamicallydetermined quantity is the Riemannian metrichij on the spatial 3-manifolds ofconstant time. These together parametrisegµν as follows:

    ds2 = −α2 (dx0)2 + hik (dxi + βidx0)(dxk + βkdx0) . (28)

    The physical interpretation ofα andβi is as follows: think of spacetime as thehistory of space. Each ‘moment’ of time,x0 = t, corresponds to an entire 3-dimensional sliceΣt. Obviously there is plenty of freedom how to ‘waft’ spacethrough spacetime. This freedom corresponds precisely to the freedom to choosethe 1+3 functionsα andβi. For one thing, you may freely specify how far for eachparameter stepdt you push space in a perpendicular direction forward in time.Thisis controlled byα, which is just the ratiods/dt of theproperperpendicular distancebetween the hypersurfacesΣt andΣt+dt. This speed may be chosen in a space andtime dependent fashion, which makesα a function on spacetime. Second, givena point with coordinatesxi on Σt. Going fromxi in a perpendicular directionyou meetΣt+dt in a point with coordinatesxi + dxi, wheredxi can be chosenat will. This freedom of moving the coordinate system aroundwhile evolving iscaptured byβi; one writesdxi = βidt. Clearly this moving around of the spatial

    10

  • coordinates can also be made in a space and time dependent fashion, so that theβi

    are functions of spacetime, too.Letnµ be the vector field in spacetime which is normal to the spatialsections of

    constant time. It is given byn = 1α(∂/∂x0 − βi∂/∂xi), as one may readily verify

    using (28) (you have to check thatn is normalized and satisfiesg(n, ∂/∂xi) = 0).We define theextrinsic curvature, Kij , to be one-half the Lie derivative of thespatial metric in the direction of the normal:

    Kij :=12Lnhij =

    1

    (

    ∂hij∂x0

    − 2D(iβj))

    , (29)

    whereD is the spatial covariant derivative with respect to the metric hij . As usual,a round bracket around indices denotes their symmetrization. Note that, by defini-tion,Kij is symmetric. Finally we denote the Ricci scalar ofhij byR(3).

    We can now write down the four constraints of the vacuum Einstein equationsin terms of these variables:

    0 = G(n, n) = 12(R(3) +KijKij − (Kii )2) , (30)

    0 = G(n, ∂/∂xj) = Di(Kij − δijKkk ) . (31)

    (30) and (31) are referred to asHamiltonian constraintandmomentum constraintrespectively. The six evolution equations of second order in the time derivativecan now be written as twelve equations of first order. Six of them are just (29),read as equation that relates the time derivative∂hij/∂x0 to the ‘canonical data’(hij ,Kij). The other six equations, whose explicit form needs not concern us here(see e.g. [22]), express the time derivative ofKij in terms of the canonical data.Both sets of evolution equations contain on their right handsides the lapse andshift functions, whose evolution is not determined but mustbe specified by hand.This specification is a choice of gauge, without which one cannot determine theevolution of the physical variables(hij ,Kij).

    The initial-data problem takes now the following form

    I. Choose a topological 3-manifoldΣ.

    II. Find onΣ a Riemannian metrichij and a symmetric tensor fieldKij whichsatisfy the constraints (30) and (31).

    III. Choose a lapse functionα and a shift vector fieldβi, both as functions ofspace and time, possibly according to some convenient prescription (e.g.singularity avoiding gauges, like maximal slicing).

    11

  • IV. Evolve initial data with these choices ofα andβi according to the twelveequations of first order. By consistency of Einstein’s equations the con-straints will be preserved during this evolution, independent of the choicesfor α andβi.

    The backbone of this setup is a mathematical theorem, which states that for any setof initial data, taken from a suitable function space, thereis, up to diffeomorphism,a unique maximal Einstein spacetime developing from these data [7].

    5 Black hole data

    5.1 Horizons

    By black hole data we understand vacuum data which containapparent horizons.The informal definition of an apparent horizon is that it is the boundary of a trappedregion, which means that its orthogonal outgoing null rays must have zero diver-gence. (Inside the trapped region they converge for any 2-surface, by definition of‘trapped region’.) The Penrose-Hawking singularity theorems state that the exis-tence of an apparent horizon implies that the evolving spacetime will be singular(assuming the strong energy-condition). Given also the condition that singularitiescannot be seen by observers far off, a condition usually calledcosmic censorship,one infers the existence of anevent horizonand hence a black hole. One can thenshow that the intersection of the event horizon with the spatial hypersurface lieson or outside the apparent horizon (for stationary spacetimes they coincide). Thereason why one does not deal with event horizons directly is that you cannot tellwhether there exists one by just looking at initial data. In principle you would haveto evolve them to the infinite future, which is beyond our abilities in general. Incontrast, apparent horizons can be recognized once the dataon an initial slice aregiven. The formal definition of an apparent horizon is the following: given initialdata(Σ, hij ,Kij) and an embedded 2-surfaceσ ⊂ Σ with outward pointing nor-mal νi. σ is an apparent horizon if and only if the following relation betweenKij,the extrinsic curvature ofΣ in spacetime, andkij, the extrinsic curvature ofσ in Σ,is satisfied:

    qijkij = −qijKij , (32)whereqij := hij − νiνj is just the induced Riemannian metric onσ, so that (32)simply says that the restriction ofKij to the tangent space ofσ has opposite traceto kij . (The minus sign on the rhs of (32) signifies afutureapparent horizon cor-responding to ablack hole which has afuture event horizon. A plus sign wouldsignify a past apparent horizon corresponding to a ‘white hole’ with past event

    12

  • horizon.) This means that once we have the data(Σ, hij ,Kij) we can in princi-ple find all 2-surfacesσ ⊂ Σ for which (32) holds and therefore find all apparenthorizons.

    5.2 Poincaŕe charges

    By Poincaré charges we shall understand quantities like mass, linear momentum,and angular momentum. In GR they are associated to an asymptotic Poincarésymmetry (see [4]), provided that the data(Σ, hij ,Kij) areasymptotically flatina suitable sense, which we now explain. Topologically asymptotic flatness meansthat the non-trivial ‘topological features’ ofΣ should all reside in a bounded regionand not ‘pile up’ at infinity. More formally this is expressedby saying that there isa bounded regionB ⊂ Σ such thatΣ−B (the complement ofB) consists of a finitenumber of disjoint pieces, each of which looks topologically like the complementof a ball inR3. These asymptotic pieces are also called theendsof the manifoldΣ. Next comes the geometric restriction imposed by the condition of asymptoticflatness. It says that for each end there is an asymptoticallyeuclidean coordinate-system{x1, x2, x3} in which the fields(hij ,Kij) have the following fall-off forr → ∞ (r =

    (x1)2 + (x2)2 + (x3)2, nk = xk/r)

    hij(xk) = δij +

    sij(nk)

    r+O(r−1−ǫ) , (33)

    Kij(xk) =

    tij(nk)

    r2+O(r−2−ǫ) . (34)

    Moreover, in order to have convergent expressions for physically relevant quanti-ties, like e.g. angular momentum (see below), the fieldsij must be anevenfunc-tion of its argument, i.e.sij(−nk) = sij(nk), andtij must be anodd function, i.e.tij(−nk) = −tij(nk).

    Under these conditions each end can be assigned mass, momentum, and angu-lar momentum, which are conserved during time evolution. They may be computedby integrals over 2-spheres in the limit the spheres are pushed to larger and largerradii into the asymptotically flat region of that end. These so-called ADM-Integrals(first considered by Arnowitt, Deser, and Misner in [1]) are given by the followingexpressions, which we give in ‘geometric’ units (meaning that in order to get themin standard units one has to multiply the mass expression given below by1/κ and

    13

  • the linear and angular momentum byc/κ):

    M = limr→∞

    S2(r)δij(∂ihjk − ∂khij)nk dσ (35)

    P i = limr→∞

    S2(r)(Kik − δikKjj )nk dσ (36)

    Si = limr→∞

    S2(r)εijlx

    j(K lk − δlkKnn )nk dσ (37)

    5.3 Maximal and time-symmetric data

    The constraints (30, 31) are too complicated to be solved in general. Further condi-tions are usually imposed to reduce the complexity of the problem: data(hij ,Kij)are calledmaximal if Kii = h

    ijKij = 0. The name derives from the fact thatKii = 0 is the necessary and sufficient condition for a hypersurfaceto have station-ary volume to first order with respect to deformations in the ambient spacetime.Even though stationarity does generally not imply extremality one calls such hy-persurfaces maximal. Note also that since spacetime is a Lorentzian manifold,extremal spacelike hypersurfaces will be of maximal ratherthan minimal volume.In contrast, in Riemannian manifolds one would speak of minimal surfaces.

    A much stronger condition is to imposeKij = 0, which as seen from (36) and(37) implies that all momenta and angular momenta vanish. Only the mass is nowallowed to be non-zero. Such data are calledtime-symmetricsince for themhijis momentarily static as seen from (29). This implies that the evolution of suchdata into the future and into the past will coincide so that the developed space-time will have a time-reversal symmetry which pointwise fixes the initial surfacewhereKij = 0. This surface is therefore also called themoment of time-symmetry.Time-symmetric data can still represent configurations of any number of blackholes without angular momenta which are momentarily at rest. Note also that fortime-symmetric data the condition (32) for an apparent horizon is equivalent tothe tracelessness of the extrinsic curvature ofσ. Hence for time symmetric dataapparent horizons are minimal surfaces.

    We add one more general comment concerning submanifolds. A vanishing ex-trinsic curvature is equivalent to the property that each geodesic of the ambientspace, which starts on, and tangent to, the submanifold, will always run entirelyinside the submanifold. Therefore, submanifolds with vanishing extrinsic curva-ture are calledtotally geodesic. Now, if the ambient space allows for an isometry(symmetry of the metric), whose fixed-point set is the submanifold in question,like for the time-reversal transformation just discussed,the submanifold must nec-essarily be totally geodesic. To see this, consider a geodesic of the ambient space

    14

  • which starts on, and tangent to, the submanifold. Assume that this geodesic even-tually leaves the submanifold. Then its image under the isometry would again be ageodesic (since isometries always map geodesics to geodesics) which is differentfrom the one we started from. But this is impossible since they share the sameinitial conditions which are known to determine the geodesic uniquely. Hencethe geodesic cannot leave the submanifold, which proves theclaim. We will laterhave more opportunities to identify totally geodesic submanifolds—namely appar-ent horizons—by their property of being fixed-point sets of isometries.

    5.4 Solution strategy for maximal data

    Possibly the most popular approach to solving the constraints is theconformaltechniquedue to York, Lichnerowicz and others (see [37] for a review).The basicidea is to regard the Hamiltonian constraint (30) as equation for the conformalfactor of the metrichij and freely specify the complementary information, calledthe conformal equivalence-class ofhij . More concretely, this works as follows:

    I. Choose unphysical (‘hatted’) quantities(ĥij , K̂ij), whereĥij is a Rieman-nian metric onΣ andK̂ij is symmetric, trace and divergence free:

    ĥijK̂ij = 0, D̂iK̂ij = 0 , (38)

    whereD̂ is the covariant derivative with respect toĥij.

    II. Solve the (quasilinear elliptic) equation for a positive, real valued functionΦ with boundary conditionΦ(r → ∞) → 1, where∆̂ = ĥijD̂iD̂j :

    ∆̂Φ + 18K̂ijK̂ijΦ

    −7 = 0 (39)

    III. Using the solution of (39), define physical (‘unhatted’) quantities by

    hij = Φ4ĥij , (40)

    Kij = Φ−2K̂ij . (41)

    These will satisfy the constraints (30, 31)!

    5.5 Explicit time symmetric data

    Before we say a little more about maximal data we wish to present some of the mostpopular examples for time symmetric data some of which are also extensively usedin numerical simulations. Hopefully these examples let yougain some intuitionin the geometries and topologies involved and also let you anticipate the richness

    15

  • that a variable space-structure gives to the solution spaceof one of the simplestequations in physics: the Laplace equation.

    Restricting the solution strategy, outlined above, to the time-symmetric caseone first observes that forKij = 0 one hasK̂ij = 0. The momentum constraint(31) is automatically satisfied and all that remains is equation (39), which nowsimply becomes the Laplace equation for the single scalar functionΦ on the Rie-mannian manifold(Σ, ĥij).

    We now make a further simplifying assumption, namely thatĥij is in fact theflat metric. This will restrict our solutionhij to aconformally flatgeometry. It isnot obvious how severe the loss of physically interesting solutions is by restrictingto conformally flat metrics. But we will see that the latter already contain manyinteresting and relevant examples.

    So let us solve Laplace’s equation in flat space! Remember that Φ must bepositive and approach 1 at spatial infinity (asymptotic flatness). We cannot takeΣ = R3 since the only solution to the Laplace equation inR3 which asymptoticallyapproaches 1 is identically 1. We must allowΦ to blow up at some points, whichwe can then remove from the manifold. In this way we let the solution tell uswhat topology to choose in order to have an everywhere regular solution. Youmight think that just removing singular points would be rather cheating, since theresulting manifold may turn out to be incomplete, that is, can be hit by a curve afterfinite proper length (you can go ‘there’), even thoughΦ and hence the physicalmetric blows up at this point. If this were the case one definitely had to say what asolution on the completion would be. But, as a matter of fact,this cannot happenand the punctured space will turn out to be complete in the physical metric.

    5.5.1 One black hole

    The simplest solution with one puncture (atr = 0) is just

    Φ(r, θ, ϕ) = 1 +a

    r, (42)

    wherea is a constant which we soon interpret and which must be positive in orderfor Φ to be positive everywhere. We cannot have other multipole contributionssince they inevitably would forceΦ to be negative somewhere. What is the geom-etry of this solution? The physical metric is

    ds2 =(

    1 +a

    r

    )4(dr2 + r2 dθ2 + sin2 θ dϕ2) (43)

    which is easily checked to be invariant under the inversion-transformation on thespherer = a:

    r → a2

    r, θ → θ, ϕ→ ϕ . (44)

    16

  • Figure 1:One black hole

    This means that the regionr > a just looks like the regionr < a and that thespherer = a has the smallest area among all spheres of constant radius. It is aminimal surface, in fact even a totally geodesic submanifold, since it is the fixedpoint set of the isometry (44). Hence it is an apparent horizon, whose area followsfrom (43):

    A = 16π(2a)2 . (45)

    Our manifold thus corresponds to a black hole. Its mass can easily be computedfrom (35); one findsm = 2a. This manifold has two ends, one forr → ∞ andone forr → 0. They have the same geometry and hence the same ADM mass, as itmust be the case since individual and total mass clearly coincide for a single hole.

    The data just written down correspond to the ‘middle’ slice right across theKruskal (maximally extended Schwarzschild) manifold. Also, (43) is just the spa-tial part of the Schwarzschild metric in isotropic coordinates. Hence we know itsentire future development in analytic form. Already for twoholes this is not thecase anymore. Even the simplest two body problem—head-on collision— has notbeen solved analytically in GR.

    5.5.2 Two black holes

    There is an obvious generalization of (42) by allowing two ‘monopoles’ of strengtha1 anda2 at the punctures~x = ~x1 and~x = ~x2 respectively. The 3-metric then reads

    ds2 =

    (

    1 +a1

    |~x− ~x1|+

    a2|~x− ~x2|

    )4

    (dr2 + r2 dθ2 + sin2 θ dϕ2). (46)

    The manifold has now three asymptotically flat ends, one for|~x| → ∞, where theoverall ADM massM is measured, and one each for|~x − ~x1,2| → 0. To see thelatter, it is best to write the metric (46) in spherical polarcoordinates(r1, θ1, ϕ1)

    17

  • Figure 2:Two black holes well separated

    centered at~x1, and then introduce the inverted radial coordinate given byr̄1 =a21/r1. In the limit r̄1 → ∞ the metric then takes the form

    ds2 =

    (

    1 +a1(1 + a2/r12)

    r̄1+O((1/r̄1)

    2)

    )4

    (dr̄2 + r̄2 (dθ21 + sin2 θ1dϕ

    21))

    (47)wherer12 = |~x1 − ~x2|. This looks just like a one-hole metric (43). Hence, ifthe black holes are well separated (compared to their size),the two-hole geometrylooks like that depicted in figure 2. By comparison with the one-hole metric wecan immediately write down the ADM masses corresponding to the three endsr, r̄1,2 → ∞ respectively:

    M = 2(a1 + a2), m1,2 = 2a1,2(1 + χ1,2), where χ1,2 =a2,1r12

    . (48)

    Momenta and angular momenta clearly vanish (moment of time symmetry). Stillassuming well separated holes, i.e.χi = ai/r12 ≪ 1, we can calculate the bindingenergy∆E =M −m1 −m2 as function of the massesmi andr12 and get

    ∆E = −m1m2r12

    (

    1− m1 +m22r12

    +O((m1,2/r12)2)

    )

    . (49)

    The leading order is just the Newtonian expression for the binding energy of twopoint-particles with massesm1,2 at distancer12. But there are corrections to thisNewtonian form which tend to diminish the Newtonian value. Note also that (49)is still not in a good form sincer12 is not an invariantly defined geometric distancemeasure. As such one might use the lengthℓ of the shortest geodesic joining thetwo apparent horizonsS1 andS2. Unfortunately these horizons are not easy tolocate analytically and hence no closed form ofℓ(m1,m2, r12) exists which couldbe inverted to eliminater12 in favour ofℓ.

    Due to the difficulty to analytically locate the two apparenthorizons we alsocannot write down an analytic expression for their area. Butwe can give upper and

    18

  • Figure 3:Two black holes after merging

    lower bounds as follows:

    16π(2ai)2 < Ai < 16π[(2ai(1 + χi)]

    2 = 16πm2i . (50)

    The lower bound simply follows from the fact that the two-hole metric (46), ifwritten down in terms of spherical polar coordinates about any of its punctures,equals the one hole metric (43) plus a positive definite correction. The upper boundfollows from the so-called Penrose inequality in Riemannian Geometry (proven in[27]), which directly states that16πm2 ≥ A for each asymptotically flat end,wherem is the mass according to (35) andA is the area of the outermost (as seenfrom that end) minimal surface.

    If the two holes approach each other to a distance comparableto the sizes ofthe holes the geometry changes in an essential way. This is shown in figure 3. Themost important new feature is that new minimal surfaces form, in fact two [10],which both enclose the two holes. The outermost of these, as seen from the upperend, denoted byS3 in figure 3, corresponds to the apparent horizon of the newlyformed ‘compound’ black hole which contains the two old ones. For two blackholes of equal mass, i.e.a1 = a2 = a, this happens approximately for a parameterratio ofa/r12 = 0.65, which in an approximate numerical translation into the ratioof individual hole mass to geodesic separation readsm/ℓ ≈ 0.26.

    5.5.3 More than two black holes

    The method can be generalized in a straightforward manner toany numbern ofblack holes with parameters(ai, ~xi), i = 1, · · · n, for the punctures. The manifoldΣ has nown+1 ends, one for|~x| → ∞ and one for each~x→ ~xi. The expressionsfor the metric and masses are then given by the obvious generalizations of (46) and(48) respectively.

    19

  • 5.5.4 Energy bounds from Hawking’s area law

    Loosely speaking, Hawking’s area law states that the surface of a black hole cannotdecrease with time. (See [23] for a simple and complete outline of the traditionaland technically slightly restricted version and [9] for thetechnically most completeproof known today.) Let us briefly explain this statement. IfΣ is a Cauchy surface(a spacelike hypersurface in spacetime) andH the event horizon (a lightlike hy-persurface in spacetime), the two intersect in a number of components (spacelike2-manifolds), each of which we assume to be a two-sphere. Each such two-sphereis called the surface of a black hole at timeΣ. Let us pick one of them and call itB. Consider next a second Cauchy surfaceΣ′ which lies to the future ofΣ. Theoutgoing null rays ofB intersectΣ′ in a surfaceB′, and the statement is now thatthe area ofB′ is larger than or equal to the area ofB (to prove this one must assumethe strong energy condition). Note that we deliberately left open the possibility thatB′ might be a proper subset of a black hole surface at timeΣ′, namely in case theoriginal hole has merged in the meantime with another one. Ifthis does not happenB′ may be called the surface of thesameblack hole at the later timeΣ.

    Following an idea of Hawking’s [24], this can be applied to the future evolutionof multi black hole data as follows. As we already mentioned,the event horizon lieson or outside the apparent horizon. Hence the area of the ‘surface’ (as just defined)of a black hole is bounded below by the area of the corresponding apparent horizon,which in turn has the lower bound stated in (50). Suppose thatafter a long time ourconfiguration settles in an approximately stationary state, at least for some interiorregion where no gravitational radiation is emitted anymore. Since our data havezero linear and angular momentum, the final state is static and uniquely given by asingle Schwarzschild hole of some final massMfinal and corresponding surface areaAfinal = 16πM

    2final. This is a direct consequence of known black hole uniqueness

    theorems (see [26] or p. 157-186 of [25] for a summary). By thearea theoremAfinal is not less than the sum of all initial apparent horizon surface areas. Thisimmediately gives

    Mfinal ≥√

    i

    Ainitiali /16π ≥ 2

    n∑

    i

    a2i . (51)

    In passing we remark that applied to a single black hole this argument shows thatit cannot lose its mass below the valuemir :=

    Ainitial/16π, called itsirreduciblemass. Back to the multi-hole case, the total initial mass is givenby the straightfor-ward generalization of (48):

    Minitial =∑

    i

    mi = 2∑

    i

    ai(1 + χi), where χi =∑

    i 6=k

    ak|~xi − ~xk|

    . (52)

    20

  • Using these two equations we can write down a lower bound for the fractionalenergy loss into gravitational radiation:

    ∆M

    M:=

    Minitial −MfinalMinitial

    ≤ 1−

    i a2i

    i ai(1 + χi). (53)

    For a collision ofn initially widely separated (χi → 0) holes of equal mass thisbecomes

    ∆M

    M= 1−

    1/n . (54)

    For just two holes this means that at most 29% of their total rest mass can beradiated away. But this efficiency can be enhanced if the energy is distributed overa larger number of black holes.

    Another way to raise the upper bound for the efficiency is to consider spinningblack holes. For two holes the maximal value of 50% can be derived by startingwith two extremal black holes (i.e. of maximal angular momentum: J = m2 ingeometric units) which merge to form a single unspinning black hole [24].

    One can also envisage a situation where one hole participates in a scatteringprocess but does not merge. Rather it gets kicked out of the collision zone andsettles without spin (for simplicity) in an quasistationary state (for some time) farapart. The question is what fraction of energy the area theorem allows it to lose.Let this be thekth hole. Thenmfinalk ≥ 2ak = minitialk /(1 + χk). Hence

    minitialk −mfinalkminitialk

    ≤ χk1 + χk

    < 1 , (55)

    showing that an appreciable efficiency can only be obtained if the data are suchthatχk is not too close to zero. This means that thekth hole was originally not toofar from the others. This seems an unlikely process. Hence itis difficult to extractenergy from a single unspinning hole.

    For a single spinning black hole the situation is again different. Spinning itdown from an extreme state to zero angular momentum sets an upper bound forthe efficiency from the area law of 29%. This follows easily from the followingrelation between mass, irreducible mass, and angular momentum for a Kerr blackhole (see e.g. [32], formula (33.60)):

    m2 = m2ir + J2/4m2ir (56)

    SettingJ = m2 one solves formir/m =√

    1/2; hence(m − mir)/m = 1 −√

    1/2 ≈ 0.29. It can moreover be shown [8] that this limit can be (theoretically)realised by the Penrose process (compare Frolov’s lecture).

    21

  • Figure 4:Multi-Schwarzschild Figure 5:Einstein-Rosen manifold

    Needless to say that realistic processes may have far less efficiency than thistheoretical bounds from the area law alone indicate. Recentnumerical studies ofthe head-on (i.e. zero angular momentum) collision of two equal-mass black holesgive a radiated energy in units of the total energy of only10−3 [3]. With angularmomentum the efficiency is of course expected to be much better. Here recentnumerical investigations give an estimate of3× 10−2 for an inspiral of two equal-mass non-spinning black holes from the innermost stable circular orbit [2]: still along way from the theoretical upper bound.

    5.5.5 Other topologies

    Other topologies can be found which support initial data with apparent horizons.For example, instead of the ‘Schwarzschild’ manifold withn+ 1 ends forn blackholes one can find one which has just two ends for any number≥ 2 of holes andwhich has been termed the Einstein-Rosen manifold [29]. Thedifference to thedata already discussed does not primarily lie in the physicsthey represent. After alltheir different topologies are hidden behind event horizons for the outside observer,even though their interaction energies are slightly different [20][22]. However, thepoint we wish to stress here is that such data can analytically and numerically bemore convenient, despite the fact that the underlying manifold might seemtopolog-ically more complicated. The reason is that these data have more symmetries, andthat coordinate systems can be found for which these symmetries take simple ana-lytic expressions. For example, in the Einstein-Rosen manifold the upper and lowerends are isometrically related by reflections about the minimal 2-spheres in eachconnecting tube, with the fixed-point sets being the apparent horizons. Hence allthe apparent horizons can analytically be easily located inthe multi-hole Einstein-Rosen manifold, in contrast to the multi-hole Schwarzschild manifold.

    22

  • Figure 6: bipolar coordinates

    Let us briefly explain thisfor two holes of equal mass.Here one starts again fromR3

    coordinatised by spherical bi-polar coordinates. These areobtained from bi-polar coordi-nates (µ, η) in the xz planeby adding an azimuthal an-gle ϕ corresponding to a ro-tation about thez axis, justlike ordinary spherical polarcoordinates are produced fromordinary polar coordinates inthexz-plane. The coordinates(µ, η) parametrise thexz-plane according toexp(µ− iη) = (ξ+c)/(ξ−c), whereξ = z + ix andc > 0 is a constant. The lines of constantµ intersect those ofconstantη orthogonally. Both families consist of circles; those in the first familyare centered on thez-axis with radiic/ sinhµ at z = c coth µ, and those in thesecond family on thex-axis with radiic/| sin η| at |x| = c cot η.

    Following an idea of Misner [31], one can borrow the method ofimages fromelectrostatics (see. e.g. chapter 2.1 in [28]) to constructsolutionsΦ to the Laplaceequation such that the metrichij = Φ4δij has a number of reflection isometriesabout two-spheres, one for each hole. In the two hole case oneuses the two two-spheresµ = ±µ0 for someµ0 > 0, which then become the apparent horizons.Using these isometries we can take two copies of our initial manifold, excise theballs |µ| > µ0 and glue the two remaining parts ‘back to back’ along the twoboundariesµ = µ0 andµ = −µ0. The isometry-property is necessary so thatthe metric continues to be smooth across the seam. This givesan Einstein-Rosenmanifold with two tubes (or ‘bridges’, as they are sometimescalled) connectingtwo asymptotically flat regions.

    In fact, we could have just takenonecopy of the original manifold, excised theballs |µ| > µ0, andmutuallyglued together the two boundariesµ = ±µ0. Thisalso gives a smooth metric across the seam and results in a manifold known asthe Misner wormhole [30]. Metrically the Misner wormhole islocally isometricto the Einstein-Rosen manifold with two tubes (which is its ‘double cover’), buttheir topologies obviously differ. This means that for the observer outside the ap-parent horizons these two data sets are indistinguishable.This is not quite true forthe Einstein-Rosen and Schwarzschild data, which are not locally isometric. Evenwithout exploring the region inside the horizons (which anyway is rendered impos-sible by existing results on topological censorship [16]) they slightly differ in their

    23

  • Figure 7:The Misner Wormhole representing two black holes

    interaction energy and other geometric quantities, like e.g. the tidal deformation ofthe apparent horizons.

    The two parametersc andµ0 now label the two-hole configurations of equalmass. (In the Schwarzschild case the two independent parameters werea ≡ a1 =a2 andr12 ≡ |~x1 − ~x2|.) But unlike the Schwarzschild case, we can now givenot only closed analytic expressions for the total massM and individual massmin terms of the two parameters, but also for the geodesic distance of the apparenthorizonsℓ. (ℓ is used as definition for ‘instantaneous distance of the two blackholes’; for the Misner wormhole, where the two apparent horizons are identified,this corresponds to the length of the shortest geodesic winding once around thewormhole.) These read

    M = 4c

    ∞∑

    n=1

    1

    sinhnµ0, m = 2c

    ∞∑

    n=1

    n

    sinhnµ0, ℓ = 2c(1 + 2mµ0). (57)

    You might rightly wonder what ‘individual mass’ should be ifthere is no internalend associated to each black hole where the ADM-formula (35)can be applied.The answer is that there are alternative definitions of ‘quasi-local-mass’ which canbe applied even without asymptotic ends. The one we used for the expressionof m above is due to Lindquist [29] and is easy to compute in connection withthe method of images. But it is lacking a deeper mathematicalfoundation. Analternative which is mathematically better founded is due to Penrose [34], whichhowever is much harder to calculate and only applies to a limited set of situations(it agrees with the ADM mass whenever both definitions apply). Amongst themare, however, all time-symmetric conformally flat data, andfor the data above thePenrose mass has fortunately been calculated in [36]. The expression form is

    24

  • rather complicated and differs from that given above, though the difference is onlyof sixth order in an expansion in (mass/distance) [20].

    In summary, we see that the problem of setting up initial datafor two blackholes of given individual mass and given separation has no unique answer. Metri-cally as well as topologically different data sets can be found which have the sameright to be called a realization of such a configuration. For holes without associatedasymptotically flat ends no unambiguous definition for quasi-local mass exists.

    5.6 Non time-symmetric data

    According to a prescription found by Bowen and York [37], we can add linear andangular momentum within the setting of maximal data. We can still use confor-mally flat data, i.e. set̂hij = δij , on multiply puncturedR3. Then the followingtwo expressions add linear momentumP i and spin angular momentumSi to thepuncture~x = ~0:

    K̂ijP =3

    2r2

    (

    P inj + P jni − (δij − ninj)(~P · ~n))

    , (58)

    K̂ijS =3

    r3((~S × ~n)inj + (~S × ~n)jni) . (59)

    It is straightforward to check that these expressions satisfy (38) (note thatD̂i = ∂i).One can also check that these data will indeed give the proposed momenta andangular momenta at infinity (i.e. at the endr → ∞). For this one may just usethe ‘hatted’ quantities in (36) and (37), since the rescaling (41) does not influencethe leading order parts in the1/r expansion ofK, which alone contribute to theseintegrals. Linearity of all these equations inK allows us to just addKP andKSand get initial data for one black hole with given momentum~P and (spin) angularmomentum~S. Moreover, we can add any finite number of expressions of the kind(58 and 59) with parameters~Pi, ~Si based at the puncture~xi, wherei = 1, · · · , n.This then leads to a data set whose total linear and angular momentum is givenby the sum

    i~Pi and

    i~Si respectively. But one may not immediately conclude

    that the~Pi and ~Si are linear and angular momenta of the individual black holes.Rather, the latter must be calculated for the internal ends of the manifold and forthis one needs to knowΦ. The task then remains to solve (39) for the conformalfactor, with blow-ups being allowed at the given punctures.

    One interesting idea to facilitate solving (39) is to first split off the singularpart ofΦ, which blows up at the punctures{~x1, · · · , ~xn} like 1/|~x − ~xi|, from theregular reminder [6] (compare also [5]). One writes

    Φ =1

    α+ U, with

    1

    α:=

    n∑

    i=n

    ai|~x− ~xi|

    , (60)

    25

  • where theai > 0 may be freely prescribed. Inserting this into (39) gives

    ∆U + β(1 + αU)−7 = 0, with β = 18α7KijKij . (61)

    The point is now the following: for~x→ ~xi the functionα tends to zero as|~x−~xi|;henceβ, too, tends to zero as|~x − ~xi|. This means that (61) has continuous coef-ficientseverywherein R3 since the1/|~x− ~xi|6–singularity at~xi of theK-squaredterm is cancelled by multiplication withα7. (Note that this relies on using theK ’sfrom (58, 59), which possess no1/rn terms withn > 3.) This means that equa-tion (61) forU can be solved on all ofR3, without the need to excise the points{~x1, · · · , ~xn} and therefore without the need to specify ‘inner’ boundary condi-tions forU ; only the ‘outer’ boundary-conditionU(r → ∞) → 1 remains. Thissimplification seems particularly useful in numerical implementations (compare[6]).

    The total mass of our configuration isM =∑

    i 2ai. The individual masses aredetermined just as in 5.5.2, by introducing the inverted radial coordinatēri = a2i /riand reading off the coefficient of the1/2r̄i term in ther̄i → ∞ expansion. Oneeasily gets

    mi = 2ai(U(~xi) + χi) (62)

    with χi as in (52).The linear and angular momenta at, say, thekth end can also be calculated by

    using inverted coordinates, given by~̄x = (~x − ~xk)a2k/r2k. Expressed in these co-ordinates, the ‘hatted’ (unphysical) extrinsic curvaturetensor is given byJ ikJ

    jl K̂ij

    whereJ ik :=∂xi

    ∂x̄k= (a2k/r̄

    2)Rik with Rik = δ

    ik − 2nink, which is an orthogonal

    matrix. The ‘physical’ extrinsic curvature is then obtained by multiplication withΦ−2 (compare 41). Now,Φ(~̄x) = r̄

    ak(1 + mk2r̄ +O((1/r̄)

    2)) so that

    K̄ij ={

    (akr̄

    )6+ terms∝

    (akr̄

    )p}

    (RkiRljK̂kl), where p ≥ 7 . (63)

    Inserting the expression (58) for̂KP results in a1/r̄4 falloff so that the individuallinear momenta are all zeroas measured from the internal ends. One may say thatthe asymptotically flat internal ends represent the local rest frames of the blackholes. Note that these rest frames are inertial since each black hole is freely falling.InsertingKS from (59) gives a1/r̄3 falloff and an angular momentum which is just−~Sk for thekth end. (Here one uses thatRik is orientation-reversing orthogonal,hence changing the sign ofεijk, and thatRik(

    ~S × ~n)k = −(~S × ~n)i.)

    26

  • 6 Problems and recent developments

    In this final section we draw attention to some of the current problems and devel-opments, without claiming completeness.

    I. Given black hole data forn holes of fixed masses and mutual separations(whatever definitions one uses here). One would like to minimize these dataon the amount of outgoing radiation energy. Any excess over the minimalamount can be said to be ‘already contained’ initially. But so far no local (intime) criterion is known which quantifies the amount of gravitational radia-tion in an initial data set. First hints at the possibility that some (Newman-Penrose) conserved quantities could be useful here were discussed in [13].

    II. Restricting to spatially conformally flat metrics seemsto be too narrow. Ithas been shown that there are no conformally flat spatial slices in Kerr space-time which are axisymmetric and reduce to slices of constantSchwarzschildtime in the limit of vanishing angular momentum [18]. Accordingly, Bowen-York data, even for a single black hole, contain excess gravitational radiationdue to the relaxation of the individual holes to Kerr form [19]. See also [33]for an informal discussion of this and related problems. An alternative to theBowen-York data which describe two spinning black holes andwhich reduceto Kerr data if the mass of one hole goes to zero have been discussed in [12].

    III. Even for the most simple 2-hole data (Schwarzschild or Einstein-Rosen) itis not known whether the evolving spacetime will have a suitably smoothasymptotic structure at future-lightlike infinity (i.e. ‘scri-plus’). As a conse-quence, we still do not know whether we can give a rigorous mathematicalmeaning to the notion of ‘energy loss by gravitational radiation’ in this caseof the simplest head-on collision of two black holes! The difficult analyticalproblems involved are studied in the framework of the so-called ‘conformalfield equations’. See [17] (in particular section 4) for a summary and refer-ences.

    IV. We usually like to ask ‘Newtonian’ questions, like: given two black holesof individual massesm1,2 and mutual separationℓ, what is their bindingenergy? For such a question to make sense we need good concepts ofquasi-local massanddistance. But these are ambiguous concepts in GR. Differentdefinitions of ‘quasi-local mass’ and ‘distance’ amount to differences in thecalculated binding energies which can be a few10−3 times total energy atclosest encounter [20]. This is of the same order of magnitude as the totalenergy lost into gravitational radiation found in [3] for head-on (i.e. zeroangular momentum) collision of two black holes modelled with Misner data.

    27

  • 7 Appendix: Eq. (2) satisfies the energy principle

    By the ‘energy principle’ we understand the property, thatall energy of the self-gravitating system serves as source for the gravitational field. In this appendix wewish to prove that (2) indeed satisfies this principle. For the uniqueness argumentsee [21].

    Given a matter distributionρ immersed in its own gravitational potentialφ.Suppose we redistribute the matter within a bounded region of space by activelydragging it along the flow lines of a vector field~ξ which vanishes outside somebounded region. The rate of change,δρ, of the matter distribution is then deter-mined throughδρ dV = −L~ξ (ρ dV ) = −~∇ · (~ξρ) dV , whereL~ξ is the Lie deriva-tive with respect to~ξ anddV is the standard spatial volume element. Note that thelatter also needs to be differentiated along~ξ, resulting inL~ξ dV =

    ~∇· ~ξ. Hence wehaveδρ = −~∇ · (~ξρ). The rate of work done to the system during this process is

    δA =

    R3

    dV ρ~ξ · ~∇φ = −∫

    R3

    dV φ ~∇ · (~ξρ) =∫

    R3

    dV φ δρ , (64)

    where the integration by parts does not lead to surface termsdue to~ξ vanishingoutside a bounded region. Equation (64) is still completelygeneral, that is, inde-pendent of the field equation forφ. The field equation comes in when we assumethat the process of redistribution is carried out adiabatically, which means that ateach stage during the processφ satisfies its field equation with the instantaneousmatter distribution. Our claim will be proven if under the hypothesis thatφ satisfies(2) we can show thatδA = c2δMG, whereMG is defined in (5) and represents thetotal gravitating energy according to the field equation. Setting

    φ/c2 = ψ andusing the more convenient equation (3), we have

    δA =c4

    2πG

    R3

    dV ψ2δ

    [

    ∆ψ

    ψ

    ]

    =c4

    2πG

    R3

    dV [ψ∆(δψ) − (∆ψ)δψ]

    =c4

    2πGlimr→∞

    S2(r)dσ ~n ·

    [

    ψ~∇(δψ) − (~∇ψ)δψ]

    . (65)

    Now, the fall-off condition forr → ∞ imply that ~∇ψ falls off as fast as1/r2 andδψ as1/r. Hence the second term in the last line of (65) does not contribute sothat we may reverse its sign. This leads to

    δA =c4

    2πGδ lim

    r→∞

    S2(r)dσ (~n · ~∇ψ)ψ = c

    4

    4πGδ lim

    r→∞

    S2(r)dσ ~n · ~∇φ

    = c2δMG (66)

    which proves the claim.

    28

  • References

    [1] Arnowitt R and Deser S and Misner C 1961 “Coordinate invariance and en-ergy expressions in general relativity” Phys. Rev.122997-1006

    [2] Baker Jet al.2001 “Plunge waveforms from inspiralling binary black holes”Phys. Rev. Lett.87121103 (gr-qc/0102037)

    [3] Baker Jet al. 2000 “Gravitational waves from black hole collisions via aneclectic approach” Class. Quant. Grav.17 L149-L156

    [4] Beig R andÓ Murchadha N 1987 “The Poincaré group as the symmetry groupof canonical general relativity”. Ann. Phys. (NY)174463-498

    [5] Beig R 2000 “Generalized Bowen–York initial data” in: Cotsakis S andGibbons G (eds.)Mathematical and Quantum Aspects of Relativity andCosmologyLecture Notes In Physics, Vol. 537 (Berlin: Springer-Verlag)(gr-qc/0005043)

    [6] Brandt S and Brügmann B 1997 “A simple construction of initial data formultiple black holes” Phys. Rev. Lett.78 3606-3609 (gr-qc/9703066)

    [7] Choquet-Bruhat Y and Geroch R 1969 “Global aspects of theCauchy prob-lem in general relativity” Comm. Math. Phys.3 334–357

    [8] Christodoulou D 1970 “Reversible and irreversible transformations on black-hole physics” Phys. Rev. Lett.25 1596-1597

    [9] Chruściel Pet al. 2001 “The area theorem”, Annales Henri Poincaré2 109-178 (gr-qc/0001003)

    [10] Čadež A. 1974 “Apparent horizons in the two-black hole problem” Ann.Phys. (NY)83449–457

    [11] Dadhich N and Narlikar J (eds.) 1998Gravitation and Relativity: At theturn of the MilleniumProceedings of the GR-15 Conference, held at IUCAA,Pune, India, Dec. 1997; published by IUCAA.

    [12] Dain S 2001 “Initial data for two Kerr-like black holes”Phys. Rev. Lett.87121102 (gr-qc/0012023)

    [13] Dain S and Valiente-Kroon J A 2002 “Conserved quantities in a black holecollision” Class. Quant. Grav.19 811-816 (gr-qc/0105109)

    29

    http://arxiv.org/abs/gr-qc/0102037http://arxiv.org/abs/gr-qc/0005043http://arxiv.org/abs/gr-qc/9703066http://arxiv.org/abs/gr-qc/0001003http://arxiv.org/abs/gr-qc/0012023http://arxiv.org/abs/gr-qc/0105109

  • [14] Damour T 1987 “The problem of motion in Newtonian and Einsteinian grav-ity” in : Hawking S and Israel W300 Years of Gravitationpp. 128-198 (Cam-bridge: Cambridge University Press)

    [15] Ehlers Jet al. 1976 “Comments on gravitational radiation damping and en-ergy loss in binary systems” Astroph. J.208L77-L81

    [16] Friedman Jet al. 1993 “Topological censorship” Phys. Rev. Lett.71 1486-1489; Erratum-ibid.75 (1995) 1872

    [17] Friedrich H 1998 “Einstein’s equation and geometric asymptotics” in: [11]pp. 153-176 (gr-qc/9804009)

    [18] Garat A and Price R 2000 “Nonexistence of conformally flat slices of the Kerrspacetime” Phys. Rev.D61 124011 (gr-qc/0002013)

    [19] Gleiser Ret al. 1998 “Evolving the Bowen-York initial data for spinningblack holes” Phys. Rev.D57 3401-3407 (gr-qc/9710096)

    [20] Giulini D 1990 “Interaction energies for three-dimensional wormholes”Class. Quant. Grav.7 1271-1290

    [21] Giulini D 1996 “Consistently implementing the fields self-energy in Newto-nian gravity” Phys. Lett.A232 165-170 (gr-qc/9605011)

    [22] Giulini D 1998 “On the construction of time-symmetric black hole initialdata” in: Hehlet al. (1998) p 224-243

    [23] Giulini D 1998 “Is there a general area theorem for blackholes?” J. Math.Phys.39 6603-6606

    [24] Hawking S 1971 “Gravitational radiation from colliding black holes” Phys.Rev. Lett.261344-1346

    [25] Hehl F et al. (eds.) 1998Black Holes: Theory and Observation, LectureNotes in Physics Vol. 514 (Berlin: Springer-Verlag)

    [26] Heusler M 1996Black Hole Uniqueness Theorems, Cambridge Lecture Notesin Physics (Cambridge: Cambridge University Press)

    [27] Huisken G and Ilmanen T 1997 “The Riemannian Penrose Inequality” Int.Math. Res. Not.20 1045-1058

    [28] Jackson J D 1975Classical Electrodynamics(second edition) (New York:John Wiley & Sons)

    30

    http://arxiv.org/abs/gr-qc/9804009http://arxiv.org/abs/gr-qc/0002013http://arxiv.org/abs/gr-qc/9710096http://arxiv.org/abs/gr-qc/9605011

  • [29] Lindquist R W 1963 “Initial-value problem on Einstein-Rosen manifolds” J.Math. Phys.4 938-950

    [30] Misner C 1960 “Wormehole initial conditions” Phys. Rev. 1181110–1111

    [31] Misner C 1963 “The method of images in geometrostatics”Ann. Phys. (NY)24 102-117

    [32] Misner C W and Thorne K S and Wheeler J A 1973Gravitation (New York:W H Freeman & Comp.)

    [33] Pullin J 1998 “Colliding black holes: analytic insights” in [11] pp. 87-105(gr-qc/9803005)

    [34] Penrose R 1982 “Quasi-local mass and angular momentum in general relativ-ity” Proc. Roy. Soc. Lond. A38153-63

    [35] Smarr L (ed.) 1997Sources of gravitational radiation(Cambridge: Cam-bridge University Press)

    [36] Tod K.P. 1983 “Some examples of Penrose’s quasi-local mass construction”Proc. Roy. Soc. Lond. A388457-477

    [37] York J 1979 “Kinematics and dynamics of general relativity” in: Smarr L(1979) p 83-126

    31

    http://arxiv.org/abs/gr-qc/9803005

    Introduction and motivationA first step beyond Newtonian gravityConstrained evolutionary structure of Einstein's equationsThe 3+1- split and the Cauchy initial-value problemBlack hole dataHorizonsPoincaré chargesMaximal and time-symmetric dataSolution strategy for maximal dataExplicit time symmetric dataOne black holeTwo black holes More than two black holesEnergy bounds from Hawking's area lawOther topologies

    Non time-symmetric data

    Problems and recent developments Appendix: Eq. (??) satisfies the energy principle