sidebar

6 Numerical Relativity

Generating time-dependent solutions to the Einstein equations using numerical methods involves an extended list of ingredients which can be loosely summarized as follows.
  • Cast the field equations as an IBVP.
  • Choose a specific formulation that admits a well-posed IBVP, i.e., there exist suitable choices for the following ingredients that ensure well posedness.
  • Choose numerically suitable coordinate or gauge conditions.
  • Discretize the resulting set of equations.
  • Handle singularities such that they do not result in the generation of non-assigned numbers which rapidly swamp the computational domain.
  • Construct initial data that solve the Einstein constraint equations and represent a realistic snapshot of the physical system under consideration.
  • Specify suitable outer boundary conditions.
  • Fix technical aspects: mesh refinement and/or multi-domains as well as use of multiple computer processors through parallelization.
  • Apply diagnostic tools that measure GWs, BH horizons, momenta and masses, and other fields.

In this section, we will discuss state-of-the-art choices for these ingredients.

6.1 Formulations of the Einstein equations

6.1.1 The ADM equations

The Einstein equations in D dimensions describing a spacetime with cosmological constant Λ and energy-matter content Tαβ are given by
( ) R − 1Rg + Λg = 8 πT ⇔ R = 8π T − ---1--g T + --2---Λg . (39 ) αβ 2 αβ αβ αβ αβ αβ D − 2 αβ D − 2 αβ
Elegant though this tensorial form of the equations is from a mathematical point of view, it is not immediately suitable for a numerical implementation. For one thing, the character of the equations as a hyperbolic, parabolic or elliptic system is not evident. In other words, are we dealing with an initial-value or a boundary-value problem? In fact, the Einstein equations are of mixed character in this regard and represent an IBVP. Well-posedness of the IBVP then requires a suitable formulation of the evolution equations, boundary conditions and initial data. We shall discuss this particular aspect in more detail further below, but first consider the general structure of the equations. The multitude of possible ways of writing the Einstein equations are commonly referred to as formulations of the equations and a good starting point for their discussion is the canonical “3 + 1” or “(D − 1) + 1” split originally developed by Arnowitt, Deser & Misner [47*] and later reformulated by York [810, 812].

The tensorial form of the Einstein equations (39*) fully reflects the unified viewpoint of space and time; it is only through the Lorentzian signature (− ,+, ..., +) of the metric that the timelike character of one of the coordinates manifests itself.9 It turns out crucial for understanding the character of Einstein’s equations to make the distinction between spacelike and timelike coordinates more explicit.

Let us consider for this purpose a spacetime described by a manifold ℳ equipped with a metric gαβ of Lorentzian signature. We shall further assume that there exists a foliation of the spacetime in the sense that there exists a scalar function t : ℳ → ℝ with the following properties. (i) The 1-form dt associated with the function t is timelike everywhere; (ii) The hypersurfaces Σt defined by t = const are non-intersecting and ∪t∈ℝΣt = ℳ. Points inside each hypersurface Σt are labelled by spatial coordinates xI, I = 1,...,D − 1, and we refer to the coordinate system (t, xI) as adapted to the spacetime split.

Next, we define the lapse function α and shift vector β through

1 α ≡ -----, βμ ≡ (∂t)μ − αnμ, (40 ) ||dt||
where n ≡ − αdt is the timelike unit normal field. The geometrical interpretation of these quantities in terms of the timelike unit normal field nα and the coordinate basis vector ∂ t is illustrated in Figure 2*. Using the relation ⟨dt,∂t⟩ = 1 and the definition of α and β, one directly finds ⟨dt,β ⟩ = 0, so that the shift β is tangent to the hypersurfaces Σt. It measures the deviation of the coordinate vector ∂t from the normal direction n. The lapse function relates the proper time measured by an observer moving with four velocity nα to the coordinate time t: Δ τ = α Δt.
View Image
Figure 2: Illustration of two hypersurfaces of a foliation Σt. Lapse α and shift βμ are defined by the relation of the timelike unit normal field μ n and the basis vector ∂t associated with the coordinate t. Note that ⟨dt,αn ⟩ = 1 and, hence, the shift vector β is tangent to Σt.

A key ingredient for the spacetime split of the equations is the projection of tensors onto time and space directions. For this purpose, the space projection operator is defined as ⊥α μ ≡ δαμ + nαn μ. For a generic tensor Tα1α2...β1β2..., its spatial projection is given by projecting each index speparately

α1α2... α1 α2 ν1 ν2 μ1μ2... (⊥T ) β1β2... ≡ ⊥ μ1⊥ μ2 ...⊥ β1⊥ β2 ...T ν1ν2.... (41 )
A tensor S is called tangent to Σt if it is invariant under projection, i.e., ⊥S = S. In adapted coordinates, we can ignore the time components of such spatial tensors and it is common practice to denote their components with Latin indices I,J, ...= 1, ...,(D − 1). We similarly obtain time projections of a tensor by contracting its indices with nα. Mixed projections are obtained by contracting any combination of tensor indices with nα and projecting the remaining ones with ⊥ αμ. A particularly important tensor is obtained from the spatial projection of the spacetime metric
γαβ ≡ (⊥g )αβ = ⊥ μα⊥ νβgμν = (δμα + n μnα)(δνβ + nνnβ)gμν = gα β + n αnβ = ⊥ αβ. (42 )
γ αβ is known as the first fundamental form or spatial metric and describes the intrinsic geometry of the spatial hypersurfaces Σt. As we see from Eq. (42*), it is identical to the projection operator. In the remainder, we will use both the ⊥ and γ symbols to denote this tensor depending on whether the emphasis is on the projection or the hypersurface geometry.

With our definitions, it is straightforward to show that the spacetime metric in adapted coordinates I (t,x ) can be written as 2 2 2 I I J J ds = − α dt + γIJ(dx + β dt)( dx + β dt) or, equivalently,

( | ) ( | ) − α2 + βM βM |βJ − α−2 | α −2βJ gαβ = -------------|--- ⇔ g αβ = --−2--I|IJ----−-2-I-J- . (43 ) βI γIJ α β γ − α β β
It can be shown [364*] that the spatial metric γIJ defines a unique, torsion-free and metric-compatible connection I 1 IM ΓJK = 2γ (∂JγKM + ∂K γMJ − ∂M γJK) whose covariant derivative for an arbitrary spatial tensor is given by
D γSα1α2...β1β2...= ⊥ λγ⊥ α1μ1⊥α2μ2 ...⊥ν1β1⊥ ν2β2 ...∇ λS μ1μ2...ν1ν2.... (44 )

The final ingredient required for the spacetime split of the Einstein equations is the extrinsic curvature or second fundamental form defined as

K αβ ≡ − ⊥∇ βn α. (45 )
The sign convention employed here is common in NR but the “−” is sometimes omitted in other studies of GR. The definition (45*) provides an intuitive geometric interpretation of the extrinsic curvature as the change in direction of the timelike unit normal field n as we move across the hypersurface Σt. As indicated by its name, the extrinsic curvature thus describes the embedding of Σt inside the higher-dimensional spacetime manifold. The projection ⊥∇ βn α is symmetric under exchange of its indices in contrast to its non-projected counterpart ∇ βn α. For the formulation of the Einstein equations in the spacetime split, it is helpful to introduce the vector field μ μ μ μ m ≡ αn = (∂t) − β. A straightforward calculation shows that the extrinsic curvature can be expressed in terms of the Lie derivative of the spatial metric along either n or m according to
1 1 K αβ = − --ℒnγ αβ = − ---ℒm γαβ. (46 ) 2 2α

We have now assembled all tools to calculate the spacetime projections of the Riemann tensor. In the following order, these are known as the Gauss, the contracted Gauss, the scalar Gauss, the Codazzi, the contracted Codazzi equation, as well as the final projection of the Riemann tensor and its contractions:

⊥ μ ⊥ ν ⊥ γ ⊥ σ Rρ = ℛ γ + K γ K − K γ K , α β ρ δ σμν δαβ α δβ β δα ⊥ μα⊥ νβR μν + ⊥ μα⊥ νβnρn σRμ ρνσ = ℛ αβ + KK αβ − K μβK αμ, R + 2R nμnν = ℛ + K2 − K μνK , γ σ μ νμνρ γ γ μν ⊥ ρn ⊥ α⊥ βR σμν = D βK α − D αK β, n σ⊥ νβRσν = D βK − D μK μβ, ⊥ αμ⊥ νβnσn ρRμ ρνσ = 1ℒmK α β + K αμK μβ + 1D αD βα, α α μ ν 1- μ 1- ⊥ α⊥ βRμν = − α ℒmK αβ − 2K αμK β − αD αD βα + ℛ αβ + KK αβ, 2 2 R = − --ℒmK − --γμνD μD να + ℛ + K2 + K μνK μν. (47 ) α α
Here, ℛ denotes the Riemann tensor and its contractions as defined in standard fashion from the spatial metric γIJ. For simplicity, we have kept all spacetime indices here even for spatial tensors. As mentioned above, the time components can and will be discarded eventually.

By using Eq. (47*), we can express the space and time projections of the Einstein equations (39*) exclusively in terms of the first and second fundamental forms and their derivatives. It turns out helpful for this purpose to introduce the corresponding projections of the energy-momentum tensor which are given by

μ ν ν μ ρ = T μνn n , jα = − ⊥ αT μνn , (48 ) Sαβ = ⊥ μα⊥ νβTμν, S = γ μνSμν. (49 )
Then, the energy-momentum tensor is reconstructed according to Tαβ = Sαβ + nαjβ + nβjα + ρn αnβ. Using the explicit expressions for the Lie derivatives
M M M ℒmKIJ = ℒ∂t−βKIJ = ∂tKIJ − β ∂M KIJ − KMJ ∂Iβ − KIM ∂J β , (50 ) ℒ γ = ℒ γ = ∂ γ − βM ∂ γ − γ ∂ βM − γ ∂ βM, (51 ) m IJ ∂t− β IJ t IJ M IJ MJ I IM J
we obtain the spacetime split of the Einstein equations
∂ γ = βM ∂ γ + γ ∂ βM + γ ∂ βM − 2αK , (52 ) t IJ M M IJ MJ I M IM J M IJ M ∂tKIJ = β ∂MKIJ + KMJ ∂Iβ + KIM ∂Jβ − DIDJ α + α (ℛIJ + KKIJ − 2KIM K J) ( S − ρ ) 2 +8π α ------γIJ − SIJ − ------α ΛγIJ, (53 ) D − 2 D − 2 0 = ℛ + K2 − KMN KMN − 2 Λ − 16πρ, (54 ) M 0 = DIK − DM K I + 8πjI. (55 )
By virtue of the Bianchi identities, the constraints (54*) and (55*) are preserved under the evolution equations. Furthermore, we can see that D (D − 1)∕2 second-order-in-time evolution equations for the γ IJ are written as a first-order-in-time system through introduction of the extrinsic curvature. Additionally, we have obtained D constraint equations, the Hamiltonian and momentum constraints, which relate data within a hypersurface Σt. We note that the Einstein equations do not determine the lapse α and shift βI. For the case of D = 4, these equations are often referred to as the ADM equations, although we note that Arnowitt, Deser & Misner used the canonical momentum in place of the extrinsic curvature in their original work [47*]. Counting the degrees of freedom, we start with D (D + 1)∕2 components of the spacetime metric. The Hamiltonian and momentum constraints determine D of these while D gauge functions represent the gauge freedom, leaving D (D − 3)∕2 physical degrees of freedom as expected.

6.1.2 Well-posedness

The suitability of a given system of differential equations for a numerical time evolution critically depends on a continuous dependency of the solution on the initial data. This aspect is referred to as well posedness of the IBVP and is discussed in great detail in Living Reviews articles and other works [645*, 674*, 383, 427]. Here, we merely list the basic concepts and refer the interested reader to these articles.

Consider for simplicity an initial-value problem in one space and one time dimension for a single variable u (t,x) on an unbounded domain. Well-posedness requires a norm || ⋅ ||, i.e., a map from the space of functions f(x) to the real numbers ℝ, and a function F (t) independent of the initial data such that

||δu(t,⋅)|| ≤ F(t)||δu (0,⋅)||, (56 )
where δu denotes a linear perturbation relative to a solution u0(t,x) [380*]. We note that F (t) may be a rapidly growing function, for example an exponential, so that well posedness represents a necessary but not sufficient criterion for suitability of a numerical scheme.

Well posedness of formulations of the Einstein equations is typically studied in terms of the hyperbolicity properties of the system in question. Hyperbolicity of a system of PDEs is often defined in terms of the principal part, that is, the terms of the PDE which contain the highest-order derivatives. We consider for simplicity a quasilinear first-order system for a set of variables u(t,x)

∂tu = P (t,x, u,∂x)u. (57 )
The system is called strongly hyperbolic if P is a smooth differential operator and its associated principal symbol is symmetrizeable [567*]. For the special case of constant coefficient systems this definition simplifies to the requirement that the principal symbol has only imaginary eigenvalues and a complete set of linearly independent eigenvectors. If linear independence of the eigenvectors is not satisfied, the system is called weakly hyperbolic. For more complex systems of equations, strong and weak hyperbolicity can be defined in a more general fashion [645, 567*, 646, 674*].

In our context, it is of particular importance that strong hyperbolicity is a necessary condition for a well posed IBVP [741, 742]. The ADM equations (52*) – (53*), in contrast, have been shown to be weakly but not strongly hyperbolic for fixed gauge [567]; likewise, a first-order reduction of the ADM equations has been shown to be weakly hyperbolic [468]. These results strongly indicate that the ADM formulation is not suitable for numerical evolutions of generic spacetimes.

A modification of the ADM equations which has been used with great success in NR is the BSSN system [78, 695] which is the subject of the next section.

6.1.3 The BSSN equations

It is interesting to note that the BSSN formulation had been developed in the 1990s before a comprehensive understanding of the hyperbolicity properties of the Einstein equations had been obtained; it was only about a decade after its first numerical application that strong hyperbolicity of the BSSN system [380] was demonstrated. We see here an example of how powerful a largely empirical approach can be in the derivation of successful numerical methods. And yet, our understanding of the mathematical properties is of more than academic interest as we shall see in Section 6.1.5 below when we discuss recent investigations of potential improvements of the BSSN system.

The modification of the ADM equations which results in the BSSN formulation consists of a trace split of the extrinsic curvature, a conformal decomposition of the spatial metric and of the traceless part of the extrinsic curvature and the introduction of the contracted Christoffel symbols as independent variables. For generality, we will again write the definitions of the variables and the equations for the case of an arbitrary number D of spacetime dimensions. We define

χ = γ−1∕(D−1), K = γMN K , MN &tidle;γIJ = χ γIJ ⇔ &tidle;γIJ = -1γIJ, χ ( ) 1 ( 1 ) &tidle;AIJ = χ KIJ − D1−1γIJK ⇔ KIJ = -- &tidle;AIJ + ------&tidle;γIJK , χ D − 1 &tidle;Γ I = &tidle;γMN &tidle;Γ IMN , (58 )
where γ ≡ det γIJ and &tidle;I ΓMN is the Christoffel symbol defined in the usual manner in terms of the conformal metric &tidle;γIJ. Note that the definition (58*) implies two algebraic and one differential constraints
&tidle;γ = 1, &tidle;γMN A&tidle; = 0, 𝒢I = Γ&tidle;I − &tidle;γMN &tidle;Γ I = 0. (59 ) MN MN

Inserting the definition (58*) into the ADM equations (52*) – (53*) and using the Hamiltonian and momentum constraints respectively in the evolution equations for K and &tidle;I Γ results in the BSSN evolution system

M --2--- M ∂tχ = β ∂M χ + D − 1 χ(αK − ∂M β ), (60 ) 2 ∂t&tidle;γIJ = βM∂M &tidle;γIJ + 2&tidle;γM (I∂J)βM − ------&tidle;γIJ∂MβM − 2α &tidle;AIJ, (61 ) D − 1 M MN &tidle;MN &tidle; --1--- 2 ∂tK = β ∂M K − χγ&tidle; DM DN α + α A AMN + D − 1 αK 8π 2 + ------α[S + (D − 3 )ρ ] −------α Λ, (62 ) D − 2 D − 2 ∂ A&tidle; = βM∂ A&tidle; + 2A&tidle; ∂ βM − --2---&tidle;A ∂ βM + αK A&tidle; − 2α &tidle;A &tidle;AM t IJ M IJ M (I J) D − 1 IJ M IJ IM J +χ (αℛ − D D α − 8παS )TF , (63 ) IJ I J IJ ∂ &tidle;Γ I = βM∂ &tidle;Γ I +---2--&tidle;Γ I∂ βM − &tidle;Γ M ∂ βI + &tidle;γMN ∂ ∂ βI + D-−-3-&tidle;γIM∂ ∂ βN t M D − 1 M M M N D − 1 M N [ ∂M χ ] D − 2 α −A&tidle;IM (D − 1 )α -----+ 2∂M α + 2 α&tidle;Γ IMN &tidle;AMN − 2------α &tidle;γIM∂M K − 16π -jI. (64 ) χ D − 1 χ
Here, the superscript “TF” denotes the trace-free part and we further use the following expressions that relate physical to conformal variables:
1 Γ IJK = &tidle;Γ IJK −--(δIK∂J χ + δIJ∂Kχ − &tidle;γJK &tidle;γIM∂M χ) , (65 ) 2χ ℛIJ = ℛ&tidle;IJ + ℛ χIJ, (66 ) [ ] ( ) ℛ χIJ = &tidle;γIJ &tidle;γMN D&tidle;M D&tidle;N χ − D-−--1&tidle;γMN ∂M χ∂N χ + D--−-3 D&tidle;I D&tidle;J χ − -1-∂Iχ ∂Jχ , (67 ) 2 χ 2χ 2χ 2 χ 1 MN M M MN [ K K ] ℛ&tidle;IJ = − -&tidle;γ ∂N ∂Nγ&tidle;IJ + &tidle;γM (I∂J)&tidle;Γ + Γ&tidle; &tidle;Γ(IJ)M + &tidle;γ 2&tidle;ΓM(I&tidle;ΓJ)KN + &tidle;ΓIM &tidle;ΓKJN , (68 ) 2 1 1 DIDJ α = D&tidle;I &tidle;DJ α + --∂(Iχ ∂J)α − ---&tidle;γIJγ&tidle;MN ∂Mχ ∂Nα. (69 ) χ 2χ
In practical applications, it turns out necessary for numerical stability to enforce the algebraic constraint MN &tidle; &tidle;γ AMN = 0 whereas enforcement of the unit determinant &tidle;γ = 1 appears to be optional. A further subtlety is concerned with the presence of the conformal connection functions &tidle;Γ I on the right-hand side of the BSSN equations. Two recipes have been identified that provide long-term stable numerical evolutions. (i) The independently evolved &tidle;I Γ are only used when they appear in differentiated form but are replaced by their definition in terms of the conformal metric &tidle;γIJ everywhere else [23*]. (ii) Alternatively, one can add to the right-hand side of Eq. (64*) a term − σ 𝒢I∂M βM, where σ is a positive constant [803].

We finally note that in place of the variable χ, alternative choices for evolving the conformal factor are in use in some NR codes, namely ϕ ≡ − (ln χ)∕4 [65*] or √ -- W ≡ χ [540]. An overview of the specific choices of variables and treatment of the BSSN constraints for the present generation of codes is given in Section 4 of [429*].

6.1.4 The generalized harmonic gauge formulation

It has been realized a long time ago that the Einstein equations have a mathematically appealing form if one imposes the harmonic gauge condition □x α = − gμνΓ αμν = 0 [294]. Taking the derivative of this condition eliminates a specific combination of second derivatives from the Ricci tensor such that its principal part is that of the scalar wave operator
1 R αβ = − -g μν∂ μ∂νgαβ + ..., (70 ) 2
where the dots denote terms involving at most the first derivative of the metric. In consequence of this simplification of the principal part, the Einstein equations in harmonic gauge can straightforwardly be written as a strongly hyperbolic system. This formulation even satisfies the stronger condition of symmetric hyperbolicity which is defined in terms of the existence of a conserved, positive energy [674], and harmonic coordinates have played a key part in establishing local uniqueness of the solution to the Cauchy problem in GR [327, 141, 321].

This particularly appealing property of the Ricci tensor can be maintained for arbitrary coordinates by introducing the functions [333, 343*]

α α μν α H ≡ □x = − g Γμν, (71 )
and promoting them to the status of independently evolved variables; see also [630*, 519*]. This is called the Generalized Harmonic Gauge formulation.

With this definition, it turns out convenient to consider the generalized class of equations

( ) --1--- ---2-- Rαβ − ∇ (α𝒞β) = 8π T αβ − D − 2T gαβ + D − 2 Λgαβ, (72 )
where 𝒞α ≡ H α − □x α. The addition of the term ∇ (α 𝒞β) replaces the contribution of ∇ (α □x β) to the Ricci tensor in terms of ∇ (αH β) and thus changes the principal part to that of the scalar wave operator. A solution to the Einstein equations is now obtained by solving Eq. (72*) subject to the constraint 𝒞α = 0.

The starting point for a Cauchy evolution are initial data gαβ and ∂tgαβ which satisfy the constraints 𝒞α = 0 = ∂t𝒞α. A convenient manner to construct such initial data is to compute the initial H α directly from Eq. (71*) so that 𝒞 α = 0 by construction. It can then be shown [519*] that the ADM constraints (54*), (55*) imply ∂ 𝒞μ = 0 t. By virtue of the contracted Bianchi identities, the evolution of the constraint system obeys the equation

[ ( ) ] μ μ 1 2 □ 𝒞α = − 𝒞 ∇ (μ𝒞α) − 𝒞 8 π Tμα − D--−-2T gμα + D-−--2Λg μα , (73 )
and the constraint 𝒞α = 0 is preserved under time evolution in the continuum limit.

A key addition to the GHG formalism has been devised by Gundlach et al. [377*] in the form of damping terms which prevent growth of numerical violations of the constraints 𝒞α = 0 due to discretization or roundoff errors.

Including these damping terms and using the definition (71*) to substitute higher derivatives in the Ricci tensor, the generalized Einstein equations (72*) can be written as

μν μν μ μ ν g ∂μ∂ νgαβ = − 2 ∂νgμ(α ∂β)g − 2 ∂(αH β) + 2H μΓαβ − 2Γ ναΓμβ 8πT--−-2Λ- [ μ ] − 8 πTαβ + D − 2 gαβ − 2κ 2n(α𝒞β) − λgαβn 𝒞μ , (74 )
where κ, λ are user-specified constraint-damping parameters. An alternative first-order system of the GHG formulation has been presented in Ref. [519].

6.1.5 Beyond BSSN: Improvements for future applications

The vast majority of BH evolutions in generic 4-dimensional spacetimes have been performed with the GHG and the BSSN formulations. It is interesting to note in this context the complementary nature of the two formulations’ respective strengths and weaknesses. In particular, the constraint subsystem of the BSSN equations contains a zero-speed mode [100, 379, 378] which may lead to large Hamiltonian constraint violations. The GHG system does not contain such modes and furthermore admits a simple way of controlling constraint violations in the form of damping terms [377]. Finally, the wave-equation-type principal part of the GHG system allows for the straightforward construction of constraint-preserving boundary conditions [650, 492*, 665*]. On the other hand, the BSSN formulation is remarkably robust and allows for the simulation of BH binaries over a wide range of the parameter space with little if any modifications of the gauge conditions; cf. Section 6.4. Combination of these advantages in a single system has motivated the exploration of improvements to the BSSN system and in recent years resulted in the identification of a conformal version of the Z4 system, originally developed in Refs. [113, 112, 114], as a highly promising candidate [28*, 163*, 775, 428*].

The key idea behind the Z4 system is to replace the Einstein equations with a generalized class of equations given by

G = 8πT − ∇ Z − ∇ Z + g ∇ Zμ + κ [n Z + n Z + κ g n ZM ], (75 ) αβ αβ α β β α αβ μ 1 α β β α 2 αβ M
where Zα is a vector field of constraints which is decomposed into space and time components according to Θ ≡ − nμZμ and ZI = ⊥ μIZμ. Clearly, a solution to the Einstein equations is recovered provided the constraint Z = 0 μ is satisfied. The conformal version of the Z4 system is obtained in the same manner as for the BSSN system and leads to time evolution equations for a set of variables nearly identical to the BSSN variables but augmented by the constraint variable Θ. The resulting evolution equations given in the literature vary in details, but clearly represent relatively minor modifications for existing BSSN codes [28*, 163, 428*]. Investigations have shown that the conformal Z4 system is indeed suitable for implementation of constraint preserving boundary conditions [664] and that constraint violations in simulations of gauge waves and BH and NS spacetimes are indeed smaller than those obtained for the BSSN system, in particular when constraint damping is actively enforced [28, 428*]. This behaviour also manifests itself in more accurate results for the gravitational radiation in binary inspirals [428*]. In summary, the conformal Z4 formulation is a very promising candidate for future numerical studies of BH spacetimes, including in particular the asymptotically AdS case where a rigorous control of the outer boundary is of utmost importance; cf. Section 6.6 below.

Another modification of the BSSN equations is based on the use of densitized versions of the trace of the extrinsic curvature and the lapse function as well as the traceless part of the extrinsic curvature with mixed indices [497, 795]. Some improvements in simulations of colliding BHs in higher-dimensional spacetimes have been found by careful exploration of the densitization parameter space [791*].

6.1.6 Alternative formulations

The formulations discussed in the previous subsections are based on a spacetime split of the Einstein equations. A natural alternative to such a split is given by the characteristic approach pioneered by Bondi et al. and Sachs [118*, 667*]. Here, at least one coordinate is null and thus adapted to the characteristics of the vacuum Einstein equations. For generic four-dimensional spacetimes with no symmetry assumptions, the characteristic formalism results in a natural hierarchy of two evolution equations, four hypersurface equations relating variables on hypersurfaces of constant retarded (or advanced) time, as well as three supplementary and one trivial equations. A comprehensive overview of characteristic methods in NR is given in the Living Reviews article [788*]. Although characteristic codes have been developed with great success in spacetimes with additional symmetry assumptions, evolutions of generic BH spacetimes face the problem of formation of caustics, resulting in a breakdown of the coordinate system; see [59] for a recent investigation. One possibility to avoid the problem of caustic formation is Cauchy-characteristic matching, the combination of a (D − 1) + 1 or Cauchy-type numerical scheme in the interior strong-field region with a characteristic scheme in the outer parts. In the form of Cauchy-characteristic extraction, i.e., ignoring the injection of information from the characteristic evolution into the inner Cauchy region, this approach has been used to extract GWs with high accuracy from numerical simulations of compact objects [642*, 60*].

All the Cauchy and characteristic or combined approaches we have discussed so far, evolve the physical spacetime, i.e., a manifold with metric (ℳ, gαβ). An alternative approach for asymptotically flat spacetimes dating back to Hübner [444] instead considers the numerical construction of a conformal spacetime &tidle; (ℳ, &tidle;gαβ) where 2 &tidle;gαβ = Ω gα β subject to the condition that gαβ satisfies the Einstein equations on ℳ. The conformal factor Ω vanishes at null infinity + − ℐ = ℐ ∪ ℐ of the physical spacetime which is thus conformally related to an interior of the unphysical manifold ℳ&tidle;, g&tidle;αβ which extends beyond the physical manifold. A version of these conformal field equations that overcomes the singular nature of the transformed Einstein equations at ℐ has been developed by Friedrich [332, 331]. This formulation is suitable for a 3 + 1 decomposition into a symmetric hyperbolic system10 of evolution equations for an enhanced (relative to the ADM decomposition) set of variables. The additional cost resulting from the larger set of variables, however, is mitigated by the fact that these include projections of the Weyl tensor that directly encode the GW content. Even though the conformal field equations have as yet not resulted in simulations of BH systems analogous to those achieved in BSSN or GHG, their elegance in handling the entire spacetime without truncation merits further investigation. For more details about the formulation and numerical applications, we refer the reader to the above articles, Lehner’s review [509], Frauendiener’s Living Reviews article [328*] as well as [329, 26] and references therein. A brief historic overview of many formulations of the Einstein equations (including systems not discussed in this work) is given in Ref. [702]; see in particular Figures 3 and 4 therein.

We finally note that for simulations of spacetimes with high degrees of symmetry, it often turns out convenient to directly impose the symmetries on the shape of the line element rather than use one of the general formalisms discussed so far. As an example, we consider the classic study by May and White [544*, 545] of the dynamics of spherically symmetric perfect fluid stars. A four-dimensional spherically symmetric spacetime can be described in terms of the simple line element

ds2 = − a2(x,t)dt2 + b2(x,t)dx2 + R2 (x,t)dΩ22, (76 )
where 2 dΩ 2 is the line element of the 2-sphere. May and White employ Lagrangian coordinates co-moving with the fluid shells which is imposed through the form of the energy-momentum tensor T 00 = − ρ (1 + 𝜖), T11 = T 22 = T33 = P. Here, the rest-mass density ρ, internal energy 𝜖, and pressure P are functions of the radial and time coordinates. Plugging the line element (76*) into the Einstein equations (39*) with D = 4, Λ = 0 and the equations of conservation of energy-momentum μ ∇ μT α = 0, result in a set of equations for the spatial and time derivatives of the metric and matter functions amenable for a numerical treatment; cf. Section II in Ref. [544] for details.

6.1.7 Einstein’s equations extended to include fundamental fields

The addition of matter to the spacetime can, in principle, be done using the formalism just laid down11. The simplest extension of the field equations to include matter is described by the Einstein–Hilbert action (in 4-dimensional asymptotically flat spacetimes) minimally coupled to a complex, massive scalar field Φ with mass parameter μ = m ∕ℏ S S,
∫ ( ) 4 √ --- R 1 μν ∗ 1 2 ∗ S = d x − g 16π-− 2g ∂μΦ ∂ νΦ − 2μ SΦ Φ . (77 )
If we introduce a time reduction variable defined as
Π = − -1(∂t − ℒβ) Φ, (78 ) α
we recover the equations of motion and constraints (52*) – (55*) with D = 4,Λ = 0 and with energy density ρ, energy-momentum flux ji and spatial components Sij of the energy-momentum tensor given by
ρ = 1Π ∗Π + 1μ2 Φ ∗Φ + 1Di Φ∗D Φ, (79 ) 21 ∗ 2 S ∗2 i ji = 2 (Π Di Φ + ΠDi Φ ), (80 ) S = 1 (D Φ ∗D Φ + D ΦD Φ ∗) + 1γ (Π∗Π − μ2Φ ∗Φ − Dk Φ ∗D Φ ) . (81 ) ij 2 i j i j 2 ij S k
Vector fields can be handled in similar fashion, we refer the reader to Ref. [794*] for linear studies and to Refs. [595, 598, 838*, 839*] for full nonlinear evolutions.

In summary, a great deal of progress has been made in recent years concerning the well-posedness of the numerical methods used for the construction of spacetimes. We note, however, that the well-posedness of many problems beyond electrovacuum GR remains unknown at present. This includes, in particular, a wide class of alternative theories of gravity where it is not clear whether they admit well-posed IBVPs.

6.2 Higher-dimensional NR in effective “3 + 1” form

Performing numerical simulations in generic higher-dimensional spacetimes represents a major challenge for simple computational reasons. Contemporary simulations of compact objects in four spacetime dimensions require 𝒪 (100 ) cores and 𝒪 (100) Gb of memory for storage of the fields on the computational domain. In the absence of spacetime symmetries, any extra spatial dimension needs to be resolved by 𝒪 (100) grid points resulting in an increase by about two orders of magnitude in both memory requirement and computation time. In spite of the rapid advance in computer technology, present computational power is pushed to its limits with D = 5 or, at best, D = 6 spacetime dimensions. For these reasons, as well as the fact that the community already has robust codes available in D = 4 dimensions, NR applications to higher-dimensional spacetimes have so far focussed on symmetric spacetimes that allow for a reduction to an effectively four-dimensional formalism. Even though this implies a reduced class of spacetimes available for numerical study, many of the most important questions in higher-dimensional gravity actually fall into this class of spacetimes. In the following two subsections we will describe two different approaches to achieve such a dimensional reduction, for the cases of spacetimes with SO (D − 2) or SO (D − 3) isometry, i.e., the rotational symmetry leaving invariant D− 3 S or D−4 S, respectively (we denote with n S the n-dimensional sphere). The group SO (D − 2) is the isometry of, for instance, head-on collisions of non-rotating BHs, while the group SO (D − 3) is the isometry of non-head-on collisions of non-rotating BHs; SO (D − 3 ) is also the isometry of non-head-on collisions of rotating BHs with one nonvanishing angular momentum, generating rotations on the orbital plane (see Figure 3*). Furthermore, the SO (D − 3 ) group is the isometry of a single rotating BH, with one non-vanishing angular momentum. We remark that, in order to implement the higher-dimensional system in (modified) four-dimensional evolution codes, it is necessary to perform a 4 + (D − 4 ) splitting of the spacetime dimensions. With such splitting, the equations have a manifest SO (D − 3) symmetry, even when the actual isometry is larger.
View Image
Figure 3: D-dimensional representation of head-on collisons for spinless BHs, with isometry group SO (D − 2) (left), and non-head-on collisons for BHs spinning in the orbital plane, with isometry group SO (D − 3 ) (right). Image reproduced with permission from [841*], copyright by APS.

We shall use the following conventions for indices. As before, Greek indices α,β, ... cover all spacetime dimensions and late upper case capital Latin indices I,J, ...= 1, ...D − 1 cover the D − 1 spatial dimensions, whereas late lower case Latin indices i,j, ...= 1,2,3 cover the three spatial dimensions of the eventual computational domain. In addition, we introduce barred Greek indices ¯ ¯α, β, ...= 0, ...,3 which also include time, and early lower case Latin indices a, b, ...= 4, ...,D − 1 describing the D − 4 spatial directions associated with the rotational symmetry. Under the 4 + (D − 4) splitting of spacetime dimensions, then, the coordinates xμ decompose as x μ → (xμ¯,xa ). When explicitly stated, we shall consider instead a 3 + (D − 3) splitting, e.g., with barred Greek indices running from 0 to 2, and early lower case Latin indices running from 3 to D − 1.

6.2.1 Dimensional reduction by isometry

The idea of dimensional reduction had originally been developed by Geroch [347] for four-dimensional spacetimes possessing one Killing field as for example in the case of axisymmetry; for numerical applications see for example Refs. [535, 704, 722, 214]. The case of arbitrary spacetime dimensions and number of Killing vectors has been discussed in Refs. [210, 211].12 More recently, this idea has been used to develop a convenient formalism to perform NR simulations of BH dynamical systems in higher dimensions, with SO (D − 2) or SO (D − 3 ) isometry [841*, 797*]. Comprehensive summaries of this approach are given in Refs. [835*, 791, 792].

The starting point is the general D-dimensional spacetime metric written in coordinates adapted to the symmetry

ds2 = g dxα dxβ = (g + e2κ2g Ba Bb ) dx ¯μdx ¯ν + 2e κBa g dx¯μ dxb + g dxa dxb. (82 ) αβ ¯μ¯ν ab ¯μ ¯ν ¯μ ab ab
Here, κ and e represent a scale parameter and a coupling constant that will soon drop out and play no role in the eventual spacetime reduction. We note that the metric (82*) is fully general in the same sense as the spacetime metric in the ADM split discussed in Section 6.1.1.

The special case of a SO (D − 2 ) (SO (D − 3)) isometry admits (n + 1)n ∕2 Killing fields ξ(i) where n ≡ D − 3 (n ≡ D − 4) stands for the number of extra dimensions. For n = 2, for instance, there exist three Killing fields given in spherical coordinates by ξ (1) = ∂ϕ, ξ(2) = sin ϕ∂𝜃 + cot𝜃 cosϕ ∂ϕ, ξ(3) = cosϕ∂ 𝜃 − cot𝜃sin ϕ∂ϕ.

Killing’s equation ℒξ(i)gAB = 0 implies that

ℒ g = 0, ℒ Ba = 0, ℒ g = 0, (83 ) ξ(i) ab ξ(i) ¯μ ξ(i)¯μ¯ν
where, as discussed above, the decomposition x μ → (xμ¯,xa ) describes a 4 + (D − 4 ) splitting in the case of SO (D − 3 ) isometry, and a 3 + (D − 3) splitting in the case of SO (D − 2) isometry.

From these conditions, we draw the following conclusions: (i) 2ψ(x¯μ) gab = e Ωab, where Ωab is the metric on the Sn sphere with unit radius and ψ is a free field; (ii) g¯μ¯ν = g¯μ¯ν(x¯σ) in adapted coordinates; (iii) [ξ(i),B ¯μ] = 0. We here remark an interesting consequence of the last property. Since, for n ≥ 2, there exist no nontrivial vector fields on Sn that commute with all Killing fields, all vector fields a B ¯μ vanish; when, instead, n = 0,1 (i.e., when D = 4, or D = 5 for SO (D − 3) isometry), this conclusion can not be made. In this approach, as it has been developed up to now [841*, 797*, 796*], one restricts to the n ≥ 2 case, and it is then possible to assume Ba ¯μ ≡ 0. Eq. (82*) then reduces to the form13

2 ¯μ ¯ν 2ψ(x¯μ) a b ds = g¯μ¯ν dx dx + e Ωabdx dx . (84 )
For this reason, this approach can only be applied when D ≥ 5 in the case of SO (D − 2) isometry, and D ≥ 6 in the case of SO (D − 3) isometry.

As mentioned above, since the Einstein equations have to be implemented in a four-dimensional NR code, we eventually have to perform a 4 + (D − 4) splitting, even when the spacetime isometry is SO (D − 2). This means that the line element is (84*), with ¯ ¯α,β, ...= 0, ...,3 and a,b, ...= 4, ...,D − 1. In this case, only a subset SO (D − 3 ) ⊂ SO (D − 2) of the isometry is manifest in the line element; the residual symmetry yields an extra relation among the components g¯μ¯ν. If the isometry group is SO (D − 3), the line element is the same, but there is no extra relation.

A tedious but straightforward calculation [835] shows that the components of the D-dimensional Ricci tensor can then be written as

R = {(D − 5) − e2ψ [(D − 4 )∂ ¯μψ∂ ψ + ¯∇ ¯μ∂ ψ ]}Ω , ab ¯μ ¯μ ab R ¯μa = 0, R ¯μ¯ν = ¯R ¯μν¯− (D − 4)(¯∇ ¯ν∂¯μψ + ∂¯μψ ∂¯νψ ), [ −2ψ ¯μ μ¯ ] R = ¯R + (D − 4) (D − 5)e − 2¯∇ ∂¯μψ − (D − 3)∂ ψ ∂¯μψ , (85 )
where ¯R μ¯¯ν, ¯R and ¯∇ respectively denote the 3 + 1-dimensional Ricci tensor, Ricci scalar and covariant derivative associated with the 3 + 1 metric ¯g¯μ¯ν ≡ g¯μ¯ν. The D-dimensional vacuum Einstein equations with cosmological constant Λ can then be formulated in terms of fields on a 3 + 1-dimensional manifold
R¯ = (D − 4)(¯∇ ∂ ψ − ∂ ψ ∂ ψ ) − Λ ¯g , (86 ) 2¯μψ¯ν[ ¯μ ν¯ ¯μ ¯μ¯μ ¯ν ] ¯μ¯ν e (D − 4 )∂ ψ∂μ¯ψ + ¯∇ ∂¯μψ − Λ = (D − 5). (87 )
One important comment is in order at this stage. If we describe the three spatial dimensions in terms of Cartesian coordinates (x,y,z ), one of these is now a quasi-radial coordinate. Without loss of generality, we choose y and the computational domain is given by x,z ∈ ℝ, y ≥ 0. In consequence of the radial nature of the y direction, 2ψ e = 0 at y = 0. Numerical problems arising from this coordinate singularity can be avoided by working instead with a rescaled version of the variable e2ψ. More specifically, we also include the BSSN conformal factor e−4ϕ in the redefinition and write
e−4ϕ ζ ≡ --2--e2ψ. (88 ) y
The BSSN version of the D-dimensional vacuum Einstein equations (86*), (87*) with Λ = 0 in its dimensionally reduced form on a 3 + 1 manifold is then given by Eqs. (60*) – (64*) with the following modifications. (i) Upper-case capital indices I,J, ... are replaced with their lower case counterparts ij, ...= 1,2,3. (ii) The (D − 1) dimensional metric γIJ, Christoffel symbols I ΓJK, covariant derivative D, conformal factor χ and extrinsic curvature variables K and A&tidle;IJ are replaced by the 3 dimensional metric γij, the 3 dimensional Christoffel symbols Γ ijk, the covariant derivative D, as well as the conformal factor χ, K and A ij defined in analogy to Eq. (58*) with D = 4, i.e.
χ = γ−1∕3, K = γnmK , mn &tidle;γij = χ γij ⇔ γ&tidle;ij = -1γij, χ ( ) 1 ( 1 ) A&tidle;ij = χ Kij − 13γijK ⇔ Kij = -- &tidle;Aij + -&tidle;γijK , χ 3 &tidle;Γ i = &tidle;γmn &tidle;Γ imn. (89 )
(iii) The extra dimensions manifest themselves as quasi-matter terms given by
yy y 4π-(ρ-+-S-) = (D − 5)χ-&tidle;γ--ζ −-1-− 2D--−-7&tidle;γmn ∂ η ∂ χ − χ &tidle;Γ--+ D--−-6-χ-&tidle;γmn ∂ ζ ∂ ζ D − 4 ζ y2 4 ζ m n y 4 ζ2 m n 1 &tidle;γym ( χ ) KK K2 + --γ&tidle;mn (χ &tidle;Dm ∂nζ − ζD&tidle;m ∂nχ) + (D − 4)---- --∂mζ − ∂m χ − ----ζ − --- 2ζ y ζ ζ 3 1 &tidle;γym D − 1 ∂ χ ∂ χ ( K K )2 − -----∂m χ + ------&tidle;γmn -m----n--− (D − 5) --ζ + --- , (90 ) 2( y ) 4 [ χ ζ 3 8π χSTiFj K ζ K 1 2χ y y 1 χ -------- = − --- + --- A&tidle;ij + -- --(δ (j∂i)ζ − ζ&tidle;Γij) + --∂iχ ∂jχ − &tidle;Di∂jχ + --D&tidle;i ∂jζ D − 4 ζ 3 ( 2 yζ ) 2χ ] ζ 1 mn χ &tidle;γym χ TF + ---&tidle;γij&tidle;γ ∂n χ ∂m χ − --∂mζ − &tidle;γij----∂m χ − --2∂iζ ∂jζ (91 ) 2(χ ) ζ (y 2ζ ) 16-πji 2- y K-ζ ym &tidle; 2- Kζ- 1- 1- 2- D − 4 = y δ iζ − γ&tidle; Ami + ζ∂iK ζ − ζ χ ∂iχ + ζ∂iζ + 3∂iK ( ) − &tidle;γnm &tidle;Ami 1∂nζ − -1∂nχ . (92 ) ζ χ
Here, 2 −1 2 K ζ ≡ − (2αy ) (∂t − ℒβ)(ζy ). The evolution of the field ζ is determined by Eq. (87*) which in terms of the BSSN variables becomes
2 βy ∂tζ = βm∂m ζ − 2αK ζ − -ζ∂m βm + 2 ζ--, (93 ) 3 y m 2- m βy- 1- γ&tidle;ym- ∂tK ζ = β ∂mK ζ − 3K ζ∂m β + 2y K ζ − 3 ζ(∂t − ℒ β)K − χζ y ∂m α [ yy ym − 1&tidle;γmn ∂ α (χ ∂ ζ − ζ∂ χ ) + α (5 − D )χ ζ&tidle;γ--−--1-+ (4 − D )χ &tidle;γ---∂ ζ 2 m n n y2 y m 2D − 7 &tidle;γym 6 − D χ 2D − 7 + -------ζ----∂m χ + --------&tidle;γmn ∂mζ ∂n ζ + -------&tidle;γmn ∂mζ ∂n χ 2 y 4 ζ 4 1 − D ζ mn K2ζ 2D − 5 D − 1 2 + -------γ&tidle; ∂m χ ∂nχ + (D − 6)--- + ------- KK ζ + ------ζK 4 χ ζ ] 3 9 1 mn &tidle;Γ y + 2&tidle;γ (ζ &tidle;Dm ∂n χ − χD&tidle;m ∂nζ) + χζ-y- . (94 )
It has been demonstrated in Ref. [841*] how all terms containing factors of y in the denominator can be regularized using the symmetry properties of tensors and their derivatives across y = 0 and assuming that the spacetime does not contain a conical singularity.

6.2.2 The cartoon method

The cartoon method has originally been developed in Ref. [25] for evolving axisymmetric four-dimensional spacetimes using an effectively two-dimensional spatial grid which employs ghostzones, i.e., a small number of extra gridpoints off the computational plane required for evaluating finite differences in the third spatial direction. Integration in time, however, is performed exclusively on the two-dimensional plane whereas the ghostzones are filled in after each timestep by appropriate interpolation of the fields in the plane and subsequent rotation of the solution using the axial spacetime symmetry. A version of this method has been applied to 5-dimensional spacetimes in Ref. [820*]. For arbitrary spacetime dimensions, however, even the relatively small number of ghostzones required in every extra dimension leads to a substantial increase in the computational resources; for fourth-order finite differencing, for example, four ghostzones are required in each extra dimension resulting in an increase of the computational domain by an overall factor 5D− 4. An elegant scheme to avoid this difficulty while preserving all advantages of the cartoon method has been developed in Ref. [630*] and is sometimes referred to as the modified cartoon method. This method has been applied to D > 5 dimensions in Refs. [700*, 512, 821] and we will discuss it now in more detail.

Let us consider for illustrating this method a D-dimensional spacetime with SO (D − 3) symmetry and Cartesian coordinates x μ = (t,x, y,z,wa ), where a = 4, ...,D − 1. Without loss of generality, the coordinates are chosen such that the SO (D − 3) symmetry implies rotational symmetry in the planes spanned by each choice of two coordinates from14 (y,wa ). The goal is to obtain a formulation of the D-dimensional Einstein equations (60*) – (69*) with SO (D − 3) symmetry that can be evolved exclusively on the xyz hyperplane. The tool employed for this purpose is to use the spacetime symmetries in order to trade derivatives off the hyperplane, i.e., in the a w directions, for derivatives inside the hyperplane. Furthermore, the symmetry implies relations between the D-dimensional components of the BSSN variables.

These relations are obtained by applying a coordinate transformation from Cartesian to polar coordinates in any of the two-dimensional planes spanned by y and w, where w ≡ wa for any particular choice of a ∈ {4, ...,D − 1}

∘ -2-----2 ρ = y + w , y = ρcos φ, φ = arctan w-, w = ρ sin φ. (95 ) y
Spherical symmetry in n ≡ D − 4 dimensions implies the existence of n(n + 1)∕2 Killing vectors, one for each plane with rotational symmetry. For each Killing vector ξ, the Lie derivative of the spacetime metric vanishes. For the yw plane, in particular, the Killing vector field is ξ = ∂ φ and the Killing condition is given by the simple relation
∂φgμν = 0. (96 )
All ADM and BSSN variables are constructed from the spacetime metric and a straightforward calculation demonstrates that the Lie derivatives along ∂ φ of all these variables vanish. For D ≥ 6, we can always choose the coordinates such that for μ ⁄= φ, gμφ = 0 which implies the vanishing of the BSSN variables βφ = &tidle;γμφ = Γ&tidle;φ = 0. The case of SO (D − 3) symmetry in D = 5 dimensions is special in the same sense as already discussed in Section 6.2.1 and the vanishing of &tidle;Γ φ does not in general hold. As before, we therefore consider in D = 5 the more restricted class of SO (D − 2) isometry which implies &tidle;Γ φ = 0. Finally, the Cartesian coordinates wa can always be chosen such that the diagonal metric components are equal,
γw1w1 = γw2w2 = ... ≡ γww. (97 )
We can now exploit these properties in order to trade derivatives in the desired manner. We shall illustrate this for the second w derivative of the ww component of a symmetric (0) 2 tensor density S of weight 𝒲 which transforms under change of coordinates x μ ↔ x ˆα according to
∂x μ ∂xν ( ∂xμ ) Sˆαˆβ = 𝒥 𝒲 ---ˆα---ˆS μν, 𝒥 ≡ det --ˆα- . (98 ) ∂x ∂x β ∂x
Specifically, we consider the coordinate transformation (95*) where 𝒥 = ρ. In particular, this transformation implies
∂ρ ∂φ ∂wSww = ---∂ρSww + ---∂φSww, (99 ) ∂w ∂w
and we can substitute
( ) Sww = 𝒥 −𝒲 -∂ρ ∂ρ-Sρρ + 2∂-ρ ∂φ-Sρφ + ∂-φ ∂φ-Sφφ . (100 ) ∂w ∂w ∂w ∂w ∂w ∂w
Inserting (100*) into (99*) and setting S ρφ = 0 yields a lengthy expression involving derivatives of S ρρ and S φφ with respect to ρ and φ. The latter vanish due to symmetry and we substitute for the ρ derivatives using
( ) [ ( ) ] ∂ S = ∂y-∂ + ∂w-∂ 𝒥 𝒲 ∂y-∂y-S + 2∂y-∂w-S + ∂w-∂w-S , ρ ρρ ∂ρ y ∂ρ w ∂ρ ∂ρ yy ∂ρ ∂ρ yw ∂ρ ∂ ρ ww ( ∂y ∂w ) [ ( ∂y ∂y ∂y ∂w ∂w ∂w ) ] ∂ ρSφφ = ---∂y + ---∂w 𝒥 𝒲 ------Syy + 2--- ---Syw + --- ---Sww . (101 ) ∂ρ ∂ρ ∂φ ∂φ ∂ φ ∂φ ∂φ ∂φ
This gives a lengthy expression relating the y and w derivatives of Sww. Finally, we recall that we need these derivatives in the xyz hyperplane and therefore set w = 0. In order to obtain an expression for the second w derivative of Sww, we first differentiate the expression with respect to w and then set w = 0. The final result is given by
∂ySww Syy − Sww ∂wSww = 0, ∂w ∂wSww = ------+ 2-----2----. (102 ) y y
Note that the density weight dropped out of this calculation, so that Eq. (102*) is valid for the BSSN variables &tidle;Aμν and &tidle;γμν as well.

Applying a similar procedure to all components of scalar, vector and symmetric tensor densities gives all expressions necessary to trade derivatives off the xyz hyperplane for those inside it. We summarize the expressions recalling our notation: a late Latin index, i = 1,...,3 stands for either x, y or z whereas an early Latin index, a = 4,...,D − 1 represents any of the a w directions. For scalar, vector and tensor fields Ψ, V and T we obtain

0 = ∂ Ψ = ∂ ∂ Ψ = V a = ∂V a = ∂ ∂ V c = ∂ V i = ∂ S = ∂ ∂ S = S a i a i a b a a bc ia bc ia = ∂a∂bSic = ∂aSij = ∂i∂aSjk, ∂yΨ ∂a∂bΨ = δab----, yy b b V-- ∂aV = δ a y , ( ∂ V y Vy ) ∂i∂aV b = δba -i--- − δiy --- , y y2 ( ∂yV i V y) ∂a∂bV i = δab -----− δiy-2- , y y Sab = δabSww, Syy − Sww ∂ySww ∂a∂bScd = (δacδbd + δadδbc) ----y2----+ δabδcd---y--, ∂ S = δ Siy −-δiySww-, a ib ab y ( ∂ S − δ ∂ S S − δ S ) ∂i∂aSjb = δab -i-jy----jy-i-ww-− δiy-jy---2jy-ww- , ( y y ) ∂ySij δiySjy + δjySiy − 2δiyδjySww ∂a∂bSij = δab --y-- − -------------y2------------ . (103 )
By trading or eliminating derivatives using these relations, a numerical code can be written to evolve D-dimensional spacetimes with SO (D − 3) symmetry on a strictly three-dimensional computational grid. We finally note that y is a quasi-radial variable so that y ≥ 0.

6.3 Initial data

In Section 6.1 we have discussed different ways of casting the Einstein equations into a form suitable for numerical simulations. At the start of Section 6, we have listed a number of additional ingredients that need to be included for a complete numerical study and physical analysis of BH spacetimes. We will now discuss the main choices used in practical computations to address these remaining items, starting with the initial conditions.

As we have seen in Section 6.1, initial data to be used in time evolutions of the Einstein equations need to satisfy the Hamiltonian and momentum constraints (54*), (55*). A comprehensive overview of the approach to generate BH initial data is given by Cook’s Living Reviews article [224*]. Here we merely summarize the key concepts used in the construction of vacuum initial data, but discuss in some more detail how solutions to the constraint equations can be generated in the presence of specific matter fields that play an important role in the applications discussed in Section 7.

One obvious way to obtain constraint-satisfying initial data is to directly use analytical solutions to the Einstein equations as for example the Schwarzschild solution in D = 4 in isotropic coordinates

( ) ( ) 2 M − 2r 2 2 M 4[ 2 2 2 2 2 ] ds = − M--+-2r- dt + 1 + 2r- dr + r( d𝜃 + sin 𝜃d ϕ ) . (104 )
Naturally, the numerical evolution of an analytically known spacetime solution does not generate new physical insight. It still serves as an important way to test numerical codes and, more importantly, analytically known solutions often form the starting point to construct generalized classes of initial data whose time evolution is not known without numerical study. Classic examples of such analytic initial data are the Misner [550] and Brill–Lindquist [133] solutions describing n non-spinning BHs at the moment of time symmetry. In Cartesian coordinates, the Brill–Lindquist data generalized to arbitrary D are given by
K = 0, γ = ψ4 ∕(D −3)δ , ψ = 1 + ∑ ------------μA-------------, (105 ) IJ IJ IJ [∑D −1 K K 2](D −3)∕2 A 4 K=1(x − x 0 )
where the summations over A and K run over the number of BHs and the spatial coordinates, respectively, and μA are parameters related to the mass of the A-th BH through the surface area ΩD −2 of the (D − 2)-dimensional sphere by μA = 16πM ∕[(D − 2 )ΩD − 2]. We remark that in the case of a single BH, the Brill–Lindquist initial data (105*) reduce to the Schwarzschild spacetime in Cartesian, isotropic coordinates (see Eq. (137*) in Section 6.7.1).

A systematic way to generate solutions to the constraints describing BHs in D = 4 dimensions is based on the York–Lichnerowicz split [515, 806, 807]. This split employs a conformal spatial metric defined by 4 γij = ψ ¯γij; note that in contrast to the BSSN variable &tidle;γij, in general det ¯γij ⁄= 1. Applying a conformal traceless split to the extrinsic curvature according to

K = A + 1γ K, Aij = ψ −10 ¯Aij ⇔ A = ψ− 2 ¯A , (106 ) ij ij 3 ij ij ij
and further decomposing A¯ ij into a longitudinal and a transverse traceless part, the momentum constraints simplify significantly; see [224*] for details as well as a discussion of the alternative physical transverse-traceless split and conformal thin-sandwich decomposition [813]. The conformal thin-sandwich approach, in particular, provides a method to generate initial data for the lapse and shift which minimize the initial rate of change of the spatial metric, i.e., data in a quasi-equilibrium configuration [225, 190].

Under the further assumption of vanishing trace of the extrinsic curvature K = 0, a flat conformal metric ¯γij = fij, where fij describes a flat Euclidean space, and asymptotic flatness limr → ∞ ψ = 1, the momentum constraint admits an analytic solution known as Bowen–York data [121*]

3 [ k ] 3 ( l k l k ) A¯ij = 2r2-Pinj + Pjni − (fij − ninj)P nk + r3 𝜖kilS n nj + 𝜖kjlS n ni , (107 )
with ∘ -2----2----2 r = x + y + z, i i n = x ∕r the unit radial vector and user-specified parameters i P, i S. By calculating the momentum associated with the asymptotic translational and rotational Killing vectors ξi(k) [811], one can show that P i and Si represent the components of the total linear and angular momentum of the initial hypersurface. The linearity of the momentum constraint further allows us to superpose solutions A¯(a) ij of the type (107*) and the total linear momentum is merely obtained by summing the individual P i (a). The total angular momentum is given by the sum of the individual spins i S(a) plus an additional contribution representing the orbital angular momentum. For the generalization of Misner data, it is necessary to construct inversion-symmetric solutions of the type (107*) using the method of images [121, 224*]. Such a procedure is not required for generalizing Brill–Lindquist data where a superposition of solutions ¯(a) A ij of the type (107*) can be used directly to calculate the extrinsic curvature from Eq. (106*) and insert the resulting expressions into the vacuum Hamiltonian constraint given with the above listed simplifications by
¯ 2 1- mn −7 ∇ ψ + 8K Kmn ψ = 0, (108 )
where ¯ 2 ∇ is the Laplace operator associated with the flat metric fij. This elliptic equation is commonly solved by decomposing ψ into a Brill–Lindquist piece ∑ ψBL = Na=1 ma ∕|⃗r − ⃗ra| plus a regular piece u = ψ − ψBL, where ⃗ra denotes the location of the a-th BH and m a a parameter that determines the BH mass and is sometimes referred to as the bare mass. Brandt & Brügmann [126] have proven existence and uniqueness of 2 C regular solutions u to Eq. (108*) and the resulting puncture data are the starting point of the majority of numerical BH evolutions using the BSSN moving puncture technique. The simplest example of this type of initial data is given by Schwarzschild’s solution in isotropic coordinates where
m Kmn = 0, ψ = 1 + 2r. (109 )
In particular, this solution admits the isometry r → m2 ∕(4r) which leaves the coordinate sphere r = m ∕2 invariant, but maps the entire asymptotically flat spacetime r > m ∕2 into the interior and vice versa. The solution, therefore, consists of 2 asymptotically flat regions connected by a “throat” and spatial infinity of the far region is compactified into the single point r = 0 which is commonly referred to as the puncture. Originally, time evolutions of puncture initial data split the conformal factor, in analogy to the initial-data construction, into a singular Brill–Lindquist contribution given by the ψ in Eq. (109*) plus a deviation u that is regular everywhere; cf. Section IV B in [24*]. In this approach, the puncture locations remain fixed on the computational domain. The simulations through inspiral and merger by [159*, 65*], in contrast, evolve the entire conformal factor using gauge conditions that allow for the puncture to move across the domain and are, therefore, often referred to as “moving puncture evolutions”.

In spite of its popularity, there remain a few caveats with puncture data that have inspired explorations of alternative initial data. In particular, it has been shown that there exist no maximal, conformally flat spatial slices of the Kerr spacetime [341, 756]. Constructing puncture data of a single BH with non-zero Bowen–York parameter i S will, therefore, inevitably result in a hypersurface containing a BH plus some additional content which typically manifests itself in numerical evolutions as spurious GWs, colloquially referred to as “junk radiation”. For rotation parameters close to the limit of extremal Kerr BHs, the amount of spurious radiation rapidly increases leading to an upper limit of the dimensionless spin parameter 2 J ∕M ≈ 0.93 for conformally flat Bowen–York-type data [226, 237, 238, 527*]; BH initial data of Bowen–York type with a spin parameter above this value rapidly relax to rotating BHs with spin χ ≈ 0.93, probably through absorption of some fraction of the spurious radiation. This limit has been overcome [527, 528] by instead constructing initial data with an extended version of the conformal thin-sandwich method using superposed Kerr–Schild BHs [467]. In an alternative approach, most of the above outlined puncture method is applied but using a non-flat conformal metric; see for instance [493, 391].

In practice, puncture data are the method-of-choice for most evolutions performed with the BSSN-moving-puncture technique15 whereas GHG evolution schemes commonly start from conformal thin-sandwich data using either conformally flat or Kerr–Schild background data. Alternatively to both these approaches, initial data containing scalar fields which rapidly collapse to one or more BHs has also been employed [629*].

The constraint equations in the presence of matter become more complex. A simple procedure can however be used to yield analytic solutions to the initial data problem in the presence of minimally coupled scalar fields [588*, 586*]. Although in general the constraints (54*) – (55*) have to be solved numerically, there is a large class of analytic or semi-analytic initial data for the Einstein equations extended to include scalar fields. The construction of constraint-satisfying initial data starts from a conformal transformation of the ADM variables [224]

4 γij = ψ ¯γij, ¯γ = det¯γij = 1, (110 ) Kij = Aij + 13γijK, Aij = ψ −2A¯ij, (111 )
which can be used to re-write the constraints as
1 1 2 5 1 ij − 7 [ i ∗ 4( ∗ 2 ∗ )] ℋ = △¯ψ − 8R¯ψ − 12K ψ + 8 ¯A ¯Aijψ + π ψ D¯ Φ D¯i Φ + ψ Π Π + μ SΦ Φ , (112 ) ℳ = D¯ A¯j − 2ψ6D¯ K − 4 πψ6 (Π∗D¯ Φ + Π ¯D Φ ∗). (113 ) i j i 3 i i i
Here, △¯ = ¯γij ¯DiD¯j, ¯D and ¯R denote the conformal covariant derivative and Ricci scalar and Π is a time reduction variable defined in (78*).

Take for simplicity a single, non-rotating BH surrounded by a scalar field (more general cases are studied in Ref. [588*, 586*]). If we adopt the maximal slicing condition K = 0 and set ¯ Aij = 0,Φ = 0, then the momentum constraint is immediately satisfied, and one is left with the the Hamiltonian constraint, which for conformal flatness, i.e., ¯γij = fij reads

[ 2 ] 1--∂- 2-∂- ---1---∂-- -∂- ---1-----∂-- 5 ∗ △ flat ψ = r2∂r r ∂r + r2sin𝜃 ∂𝜃 sin 𝜃∂ 𝜃 + r2sin2𝜃 ∂Φ2 ψ = − πψ ΠΠ . (114 )
The ansatz
−5∕2 Π = ψ√----F (r)Z(𝜃,ϕ ), (115 ) rπ M ∑ u (r) ψ = 1 + --- + -lm---Ylm(𝜃,ϕ ), (116 ) 2r lm r
reduces the Hamiltonian constraint to
( ) ∑ ′′ l(l-+-1)- 2 2 u lm − r2 ulm Ylm = − F(r) Z (𝜃,ϕ) . (117 ) lm
By a judicious choice of the angular function Z (𝜃,ϕ), or in other words, by projecting Z(𝜃,ϕ ) onto spherical harmonics Ylm, the above equation reduces to a single second-order, ordinary differential equation. Thus, the complex problem of finding appropriate initial data for massive scalar fields was reduced to an almost trivial problem, which admits some interesting analytical solutions [588*, 586*]. Let us focus for definiteness on spherically symmetric solutions (we refer the reader to Ref. [588*, 586*] for the general case), by taking a Gaussian-type solution ansatz,
1 √ -- (r−r0)2 Z (𝜃,ϕ) = √----, F (r) = A00 × re− w2 , (118 ) 4π
where A00 is the scalar field amplitude and r0 and w are the location of the center of the Gaussian and its width. By solving Eq. (117*), we obtain the only non-vanishing component of u (r) lm
2 [ ( √ -- ) ] 2 u = A2 w-[w--−--4r√0(r −-r0)] erf --2(r −-r0) − 1 − A2 r0√w--e−2(r− r0)2∕w2, (119 ) 00 00 16 2 w 008 π
where we have imposed that ulm → 0 at infinity. Other solutions can be obtained by adding a constant to (119*).

6.4 Gauge conditions

We have seen in Section 6.1, that the Einstein equations do not make any predictions about the gauge functions; the ADM equations leave lapse α and shift i β unspecified and the GHG equations make no predictions about the source functions H α. Instead, these functions can be freely specified by the user and represent the coordinate or gauge-invariance of the theory of GR. Whereas the physical properties of a spacetime remain unchanged under gauge transformations, the performance of numerical evolution schemes depends sensitively on the gauge choice. It is well-known, for example, that evolutions of the Schwarzschild spacetime employing geodesic slicing α = 1 and vanishing shift i β = 0 inevitably reach a hypersurface containing the BH singularity after a coordinate time interval t = πM [709]; computers respond to singular functions with non-assigned numbers which rapidly swamp the entire computational domain and render further evolution in time practically useless. This problem can be avoided by controlling the lapse function such that the evolution in proper time slows down in the vicinity of singular points in the spacetime [312]. Such slicing conditions are called singularity avoiding and have been studied systematically in the form of the Bona-Massó family of slicing conditions [116]; see also [343, 20]. A potential problem arising from the use of singularity avoiding slicing is the different progress in proper time in different regions of the computational domain resulting in a phenomenon often referred to as “grid stretching” or “slice stretching” which can be compensated with suitable non-zero choices for the shift vector [24*].

The particular coordinate conditions used with great success in the BSSN-based moving puncture approach [159, 65] in D = 4 dimensions are variants of the “1+log” slicing and “Γ-driver” shift condition [24]

∂tα = βm ∂m α − 2αK, (120 ) ∂tβi = βm ∂m βi + 3Bi, (121 ) 4 ∂tBi = βm ∂mBi + ∂t&tidle;Γ i − ηBi. (122 )
We note that the variable i B introduced here is an auxiliary variable to write the second-order-in-time equation for the shift vector as a first-order system and has no relation with the variable of the same name introduced in Eq. (82*). The “damping” factor η in Eq. (122*) is specified either as a constant, a function depending on the coordinates xi and BH parameters [683], a function of the BSSN variables [559, 560], or evolved as an independent variable [29]. A first-order-in-time evolution equation for i β has been suggested in [758] which results from integration of Eqs. (121*), (122*)
i m i 3&tidle;i i ∂tβ = β ∂m β + 4Γ − ηβ . (123 )
Some NR codes omit the advection derivatives of the form m β ∂m in Eqs. (120*) – (123*). Long-term stable numerical simulations of BHs in higher dimensions require modifications in the coefficients in Eqs. (120*) – (123*) [700*] and/or the addition of extra terms [841*]. Reference [313] recently suggested a modification of Eq. (120*) for the lapse function α that significantly reduces noise generated by a sharp initial gauge wave pulse as it crosses mesh refinement boundaries.

BH simulations with the GHG formulation employ a wider range of coordinate conditions. For example, Pretorius’ breakthrough evolutions [629*] set Hi = 0 and

α-−--1 μ □Ht = − ξ1 αη + ξ2n ∂μHt, (124 )
with parameters ξ1 = 19∕m, ξ2 = 2.5∕m, η = 5 where m denotes the mass of a single BH. An alternative choice used with great success in long binary BH inspiral simulations [735] sets Hα such that the dynamics are minimized at early stages of the evolution, gradually changes to harmonic gauge H α = 0 during the binary inspiral and uses a damped harmonic gauge near merger
[ ( √ -) ]2[ ( √ -) ] H α = μ0 ln --γ- ln --γ- n α − α −1gαm βm , (125 ) α α
where μ0 is a free parameter. We note in this context that for D = 4, the GHG source functions α H are related to the ADM lapse and shift functions through [630*]
μ μ n H μ = − K − n ∂μ ln α, (126 ) μi mn i im 1- μ i γ H μ = − γ Γmn + γ ∂m ln α + α n ∂μβ . (127 )

6.5 Discretization of the equations

In the previous sections, we have derived formulations of the Einstein equations in the form of an IBVP. Given an initial snapshot of the physical system under consideration, the evolution equations, as for example in the form of the BSSN equations (60*) – (64*), then predict the evolution of the system in time. These evolution equations take the form of a set of nonlinear partial differential equations which relate a number of grid variables and their time and spatial derivatives. Computers, on the other hand, exclusively operate with (large sets of) numbers and for a numerical simulation we need to translate the differential equations into expressions relating arrays of numbers.

The common methods to implement this discretization of the equations are finite differencing, the finite element, finite volume and spectral methods. Finite element and volume methods are popular choices in various computational applications, but have as yet not been applied to time evolutions of BH spacetimes. Spectral methods provide a particularly efficient and accurate approach for numerical modelling provided the functions do not develop discontinuities. Even though BH spacetimes contain singularities, the use of singularity excision provides a tool to remove these from the computational domain. This approach has been used with great success in the SpEC code to evolve inspiralling and merging BH binaries with very high accuracy; see, e.g., [122, 220, 526]. Spectral methods have also been used successfully for the modelling of spacetimes with high degrees of symmetry [205, 206, 207*] and play an important role in the construction of initial data [39*, 38, 836*]. An indepth discussion of spectral methods is given in the Living Reviews article [365]. The main advantage of finite differencing methods is their comparative simplicity. Furthermore, they have proved very robust in the modelling of rather extreme BH configurations as for example BHs colliding near the speed of light [719*, 587*, 716*] or binaries with mass ratios up to 1 : 100 [525, 523, 718].

Mesh refinement and domain decomposition:
BH spacetimes often involve lengthscales that differ by orders of magnitude. The BH horizon extends over lengths of the order 𝒪 (1 )M where M is the mass of the BH. Inspiralling BH binaries, on the other hand, emit GWs with wavelengths of 2 𝒪 (10 )M. Furthermore, GWs are rigorously defined only at infinity. In practice, wave extraction is often performed at finite radii but these need to be large enough to ensure that systematic errors are small. In order to accomodate accurate wave extraction, computational domains used for the modelling of asymptotically flat BH spacetimes typically have a size of 3 𝒪 (10 )M. With present computational infrastructure it is not possible to evolve such large domains with a uniform, high resolution that is sufficient to accurately model the steep profiles arising near the BH horizon. The solution to this difficulty is the use of mesh refinement, i.e., a grid resolution that depends on the location in space and may also vary in time. The use of mesh refinement in BH modelling is simplified by the remarkably rigid nature of BHs which rarely exhibit complicated structure beyond some mild deformation of a sphere. The requirements of increased resolution are, therefore, simpler to implement than, say, in the modelling of airplanes or helicopters. In BH spacetimes the grid resolution must be highest near the BH horizon and it decreases gradually at larger and larger distances from the BH. In terms of the internal book-keeping, this allows for a particularly efficient manner to arrange regions of refinement which is often referred to as moving boxes. A set of nested boxes with outwardly decreasing resolution is centered on each BH of the spacetime and follows the BH motion. These sets of boxes are immersed in one or more common boxes which are large enough to accomodate those centered on the BHs. As the BHs approach each other, boxes originally centered on the BHs merge into one and become part of the common-box hierarchy. A snapshot of such moving boxes is displayed in Figure 4*.

View Image
Figure 4: Illustration of mesh refinement for a BH binary with one spatial dimension suppressed. Around each BH (marked by the spherical AH), two nested boxes are visible. These are immersed within one large, common grid or refinement level.

Mesh refinement in NR has been pioneered by Choptuik in his seminal study on critical phenomena in the collapse of scalar fields [212*]. The first application of mesh refinement to the evolution of BH binaries was performed by Brügmann [140*]. There exists a variety of mesh refinement packages available for use in NR including Bam [140], Had [384], Pamr/Amrd [754], Paramesh [534], Samrai [672] and the Carpet [684*, 184] package integrated into the Cactus Computational Toolkit [155]. For additional information on Cactus see also the Einstein Toolkit webpage [300] and the lecture notes [840]. A particular mesh-refinement algorithm used for many BH applications is the Berger–Oliger [90] scheme where coarse and fine levels communicate through interpolation in the form of the prolongation and restriction operation; see [684] for details. Alternatively, the different lengthscales can be handled efficiently through the use of multiple domains of different shapes. Communication between the individual subdomains is performed either through overlaps or directly at the boundary for touching domains. Details of this domain decomposition can be found in [618, 146] and references therein.

6.6 Boundary conditions

In NR, we typically encounter two types of physical boundaries, (i) inner boundaries due to the treatment of spacetime singularities in BH solutions and (ii) the outer boundary either at infinite distance from the strong-field sources or, in the form of an approximation to this scenario, at the outer edge of the computational domain at large but finite distances.

Singularity excision:
BH spacetimes generically contain singularities, either physical singularities with a divergent Ricci scalar or coordinate singularities where the spacetime curvature is well behaved but some tensor components approach zero or inifinite values. In the case of the Schwarzschild solution in Schwarzschild coordinates, for example, r = 0 corresponds to a physical singularity whereas the singular behaviour of the metric components gtt and grr at r = 2M merely reflects the unsuitable nature of the coordinates as r → 2M and can be cured, for example, by transforming to Kruskal–Szekeres coordinates; cf. for example Chapter 7 in Ref. [186]. Both types of singularities give rise to trouble in the numerical modelling of spacetimes because computers only handle finite numbers. Some control is available in the form of gauge conditions as discussed in Section 6.4; the evolution of proper time is slowed down when the evolution gets close to a singularity. In general, however, BH singularities require some special numerical treatment.

View Image
Figure 5: Illustration of singularity excision. The small circles represent vertices of a numerical grid on a two-dimensional cross section of the computational domain containing the spacetime singularity, in this case at the origin. A finite region around the singularity but within the event horizon (large circle) is excluded from the numerical evolution (white circles). Gray circles represent the excision boundary where function values can be obtained through regular evolution in time using sideways derivative operators as appropriate (e.g., [630*]) or regular update with spectral methods (e.g., [677*, 678*]), or through extrapolation (e.g., [703*, 723*]). The regular evolution of exterior grid points (black circles) is performed with standard techniques using information also on the excision boundary.

Such a treatment is most commonly achieved in the form of singularity or BH excision originally suggested by Unruh as quoted in [746]. According to Penrose’s cosmic censorship conjecture, a spacetime singularity should be cloaked inside an event horizon and the spacetime region outside the event horizon is causally disconnected from the dynamics inside (see Section 3.2.1). The excision technique is based on the corresponding assumption that the numerical treatment of the spacetime inside the horizon has no causal effect on the exterior. In particular, excising a finite region around the singularity but within the horizon should leave the exterior spacetime unaffected. This is illustrated in Figure 5* where the excision region is represented by small white circles which are excluded from the numerical evolution. Regular grid points, represented in the figure by black circles, on the other hand are evolved normally. As we have seen in the previous section, the numerical evolution in time of functions at a particular grid point typically requires information from neighbouring grid points. The updating of variables at regular points, therefore, requires data on the excision boundary represented in Figure 5* by gray circles. Inside the BH horizon, represented by the large circle in the figure, however, information can only propagate inwards, so that the variables on the excision boundary can be obtained through use of sideways derivative operators (e.g., [630]), extrapolation (e.g., [703, 723*]) or regular update with spectral methods (e.g., [677, 678*]). Singularity excision has been used with great success in many numerical BH evolutions [23, 723, 721*, 629*, 678, 70*, 418].

Quite remarkably, the moving puncture method for evolving BHs does not employ any such specific numerical treatment near BH singularities, but instead applies the same evolution procedure for points arbitrarily close to singularities as for points far away and appears “to get away with it”. In view of the remarkable success of the moving puncture method, various authors have explored the behaviour of the puncture singularity in the case of a single Schwarzschild BH [392, 136, 137, 390, 138*, 264]. Initially, the puncture represents spatial infinity on the other side of the wormhole geometry compactified into a single point. Under numerical evolution using moving puncture gauge conditions, however, the region immediately around this singularity rapidly evolves into a so-called trumpet geometry which is partially covered by the numerical grid to an extent that depends on the numerical resolution; cf. Figure 1 in [138]. In practice, the singularity falls through the inevitably finite resolution of the computational grid which thus facilitates a natural excision of the spacetime singularity without the need of any special numerical treatment.

Outer boundary:
Most physical scenarios of interest for NR involve spatial domains of infinite extent and there arises the question how these may be accomodated inside the finite memory of a computer system. Probably the most elegant and rigorous method is to apply a spatial compactification, i.e., a coordinate transformation that maps the entire domain including spatial infinity to a finite coordinate range. Such compactification is best achieved in characteristic formulations of the Einstein equations where the spacetime foliation in terms of ingoing and/or outgoing light cones may ensure adequate resolution of in- or outgoing radiation throughout the entire domain. In principle, such a compactification can also be implemented in Cauchy-type formulations, but here it typically leads to an increasing blueshift of radiative signals as they propagate towards spatial infinity. As a consequence, any discretization method applied will eventually fail to resolve the propagating features. This approach has been used in Pretorius’ breakthrough [629] and the effective damping of radiative signals at large distances through underresolving them approximates a no-ingoing-radiation boundary condition. An intriguing alternative consists in using instead a space-time slicing of asymptotically null hypersurfaces which play a key role in the conformal field equations [328]. To our knowledge, this method has not yet been applied successfully to BH simulations in either astrophysical problems or simulations of the type reviewed here, but may well merit more study in the future. The vast majority of Cauchy-based NR applications instead resort to an approximative treatment where the infinite spatial domain is truncated and modeled as a compact domain with “suitable” outer boundary conditions. Ideally, the boundary conditions would satisfy the following requirements [651*]: They (i) ensure well posedness of the IBVP, (ii) are compatible with the constraint equations, and (iii) correctly represent the physical conditions, which in almost all practical applications means that they control or minimize the ingoing gravitational radiation.

Boundary conditions meeting these requirements at least partially have been studied most extensively for the harmonic or generalized harmonic formulation of the Einstein equations [492, 651, 58, 652*, 665].

For the BSSN system, such boundary conditions have as yet not been identified and practical applications commonly apply outgoing radiation or Sommerfeld boundary conditions, which are an approximation in this context, where they are applied at large but finite distances from the strong-field region. Let us assume, for this purpose, that a given grid variable f asymptotes to a constant background value f 0 in the limit of large r and contains a leading order deviation u(t − r)∕rn from this value, where u remains finite as r → ∞, and n is a constant positive integer number. For r → ∞, we therefore have

f(t,r) = f + u(t −-r), (128 ) 0 rn
where the dependence on retarded time represents the outgoing nature of the radiative deviations. In consequence, ∂tu + ∂ru = 0, which translates into the following conditon for the grid variable f
f − f xI ∂tf + n -----0-+ --∂If = 0, (129 ) r r
where xI denote Cartesian coordinates. Because information only propagates outwards, the spatial derivative ∂If is evaluated using a one-sided stencil. This method is straightforwardly generalized to asymptotically expanding cosmological spacetimes of dS type containing BHs; cf. Eq. (9) in Ref. [837*]. Even though this approximation appears to work rather (one might be tempted to say surprisingly) well in practice [652], it is important to bear in mind the following caveats. (i) The number of conditions imposed in this way exceeds the number of ingoing characteristics calling into question the well-posedness of the resulting system. (ii) Sommerfeld conditions are not constraint satisfying which leads to systematic errors that do not converge away as resolution is increased. (iii) Some spurious reflections of gravitational waves may occur, especially when applied at too small radii. These potential difficulties of BSSN evolutions have motivated studies of generalizing the BSSN system, in particular the conformal Z4 formulations discussed in Section 6.1.5 which accomodate constraint preserving boundary conditions which facilitate control of the ingoing gravitational radiation [428].

In asymptotically AdS spacetimes, the outer boundary represents a more challenging problem and the difficulties just discussed are likely to impact numerical simulations more severely if not handled appropriately. This is largely a consequence of the singular behaviour of the AdS metric even in the absence of a BH or any matter sources. The AdS metric (see Section 3.3.1) is the maximally symmetric solution to the Einstein equations (39*) with Tαβ = 0 and Λ < 0. This solution can be represented by the hyperboloid ∑D −1 X20 + X2D − i=1 X2i embedded in a flat D + 1-dimensional spacetime with metric ∑ ds2 = − dX20 − dX2D + D− 1dX2i . i=1 It can be represented in global coordinates, as

2 L2 2 2 2 2 ds = ---2--(− dτ + dρ + sin ρ dΩ D−2), (130 ) cos ρ
where 0 ≤ ρ < π∕2, − π < τ ≤ π and Λ = − (D − 1)(D − 2)∕(2L2) (by unwrapping the cylindrical time direction, the range of the time coordinate is often extended to τ ∈ ℝ), or in the Poincarè coordinates:
[ ] 2 D∑−2 ds2 = L-- − dt2 + dz2 + (dxi)2 , (131 ) z2 i=1
with z > 0, t ∈ ℝ. It can be shown that Poincaré coordinates only cover half the hyperboloid and that the other half corresponds to z < 0 [81].

Clearly, both the global (130*) and the Poincaré (131*) versions of the AdS metric become singular at their respective outer boundaries ρ → π∕2 or z → 0. The induced metric at infinity is therefore only defined up to a conformal rescaling. This remaining freedom manifests itself in the boundary topology of the global and Poincaré metrics which, respecitvely, become in the limit ρ → π∕2 and z → 0

2 2 2 2 2 D∑− 2 i 2 dsgl ∼ − dτ + d ΩD −2, dsP ∼ − dt + (dx ) . (132 ) i=1
In the context of the gauge/gravity duality, gravity in global or Poincaré AdS is related to CFTs on spacetimes of different topology: ℝ × SD −2 in the former and D −1 ℝ in the latter case.

The boundary treatment inside a numerical modelling of asymptotically AdS spacetimes needs to take care of the singular nature of the metric. In practice, this is achieved through some form of regularization which makes use of the fact that the singular piece of an asymptotically AdS spacetime is known in analytic form, e.g., through Eqs. (130*) or (131*). In Ref. [70*] the spacetime metric is decomposed into an analytically known AdS part plus a deviation which is regular at infinity. In this approach, particular care needs to be taken of the gauge conditions to ensure that the coordinates remain compatible with this decomposition throughout the simulation. An alternative approach consists in factoring out appropriate factors involving the bulk coordinate as for example the term cosρ in the denominator on the right-hand side of Eq. (130*). This method is employed in several recent works [207*, 415, 108*].

We finally note that the boundary plays an active role in AdS spacetimes. The visualization of the AdS spacetime in the form of a Penrose diagram demonstrates that it is not globally hyperbolic, i.e., there exists no Cauchy surface on which initial data can be specified in such a way that the entire future of the spacetime is uniquely determined. This is in marked contrast to the Minkowski spacetime. Put in other words, the outer boundary of asymptotically flat spacetimes is represented in a Penrose diagram by a null surface such that information cannot propagate from infinity into the interior spacetime. In contrast, the outer boundary of asymptotically AdS spacetimes is timelike and, hence, the outer boundary actively influences the evolution of the interior. The specification of boundary conditions in NR applications to the gauge/gravity duality or AdS/CFT correspondence therefore reflects part of the description of the physical system under study; cf. Section 7.8.

6.7 Diagnostics

Once we have numerically generated a spacetime, there still remains the question of how to extract physical information from the large chunk of numbers the computer has written to the hard drive. This analysis of the data faces two main problems in NR applications, (i) the gauge or coordinate dependence of the results and (ii) the fact that many quantities we are familiar with from Newtonian physics are hard or not even possible to define in a rigorous fashion in GR. In spite of these difficulties, a number of valuable diagnostic tools have been developed and the purpose of this section is to review how these are extracted.

The physical information is often most conveniently calculated from the ADM variables and we assume for this discussion that a numerical solution is available in the form of the ADM variables γIJ, KIJ, α and βI. Even if the time evolution has been performed using other variables as for example the BSSN or GHG variables the conversion between these and their ADM counterparts according to Eq. (58*) or (43*) is straightforward.

One evident diagnostic directly arises from the structure of the Einstein equations where the number of equations exceeds the number of free variables; cf. the discussion following Eq. (55*). Most numerical applications employ “free evolutions” where the evolution equations are used for updating the grid variables. The constraints are thus not directly used in the numerical evolution but need to be satisfied by any solution to the Einstein equations. A convergence analysis of the constraints (see for example Figure 3 in Ref. [714]) then provides an important consistency check of the simulations.

Before reviewing the extraction of physical information from a numerical simulation, we note a potential subtlety arising from the convention used for Newton’s constant in higher-dimensional spacetimes. We wrote the Einstein equations in the form (39*) and chose units where G = 1 and c = 1. This implies that the Einstein equations have the form R αβ − 1∕2Rg αβ + Λgα β = 8πGT αβ for all spacetime dimensionalities (here and in Section 6.7.1 we explicitly keep G in the equations). As we shall see below, with this convention the Schwarzschild radius of a static BH in D dimensions is given by

(D−1)∕2 D −3 --16-πGM----- 2π(----)- R S = (D − 2)ΩD − 2, ΩD −2 = Γ D−-1 , (133 ) 2
where ΩD −2 denotes the area of the unit SD −2 sphere.

6.7.1 Global quantities and horizons

For spacetimes described by a metric that is asymptotically flat and time independent, the total mass-energy and linear momentum are given by the ADM mass and ADM momentum, respectively. These quantities arise from boundary terms in the Hamiltonian of GR and were derived by Arnowitt, Deser & Misner [47] in their canonical analysis of the theory. They are given in terms of the ADM variables by
∮ ---1-- MN K MADM = 16 πG rl→im∞ δ (∂N γMK − ∂K γMN )ˆr dS, (134 ) ∮ Sr P = --1-- lim (K − δ K )ˆrM dS. (135 ) I 8 πG r→ ∞ Ωr MI MI
Here, the spatial tensor components γIJ and KIJ are assumed to be given in Cartesian coordinates, ˆrM = xM ∕r is the outgoing unit vector normal to the area element dS of the SD −2 sphere and dS = rD −2d ΩD −2. The above integral is defined only for a restricted class of coordinate systems, known as asymptotic Euclidian coordinates for which the metric components are required to be of the form gμν = ημν + 𝒪 (1∕r). Under a more restrictive set of assumptions about the fall-off behaviour of the metric and extrinsic curvature components (see Sections 7.5.1 and 7.5.2 in [364*] and references therein for a detailed discussion), one can also derive an expression for the global angular momentum
∮ 1 J K JI = 8π-lri→m∞ (KJK − K γJK)ξ(I)ˆr dS, (136 )
where ξ(I) are the Killing vector fields associated with the asymptotic rotational symmetry given, in D = 4, by ξ(x) = − z∂y + y∂z, ξ(y) = − x∂z + z ∂x and ξ(z) = − y∂x + x∂y. For a more-in-depth discussion of the ADM mass and momentum as well as the conditions required for the definition of the angular momentum the reader is referred to Section 7 of [364*]. Expressions for MADM, PI and JI can also be derived in more general (curvilinear) coordinate systems as long as the metric approaches the flat-space form in those curvilinear coordinates at an appropriate rate; see, e.g., Section 7 in [364] for a detailed review.

As an example, we calculate the ADM mass of the D-dimensional Schwarzschild BH in Cartesian, isotropic coordinates I (t,x ) described by the spatial metric

--4 μ γIJ = ψD −3δIJ, ψ = 1 + --D-−3, (137 ) 4r
and vanishing extrinsic curvature KIJ = 0. A straightforward calculation shows that
∂ γ = − ψ D4−3−1--μ--x δ , (138 ) K IJ rD −1 K IJ
so that (since xK = rˆrK)
K K δMN(∂N γMK − ∂KγMN )x-- = (D − 2)ψD4−-3−1 μxK-x--= (D − 2 )-μ--, (139 ) r rD− 1 r rD−2
where we have used the fact that in the limit r → ∞ we can raise and lower indices with the Euclidean metric δ IJ and ψ → 1. From Eq. (134*) we thus obtain
∮ ∮ --1--- -μ--- D−2 D--−-2 MADM = 16 πG rli→m∞ (D − 2 )rD−2r d ΩD −2 = 16 πG μ dΩD −2 Sr D-−1 D--−-2 D--−-2 -2π--2-- = 16πG μ ΩD −2 = 16 πG μ Γ (D−1-). (140 ) 2
The Schwarzschild radius in areal coordinates is given by RDS− 3= μ and we have recovered Eq. (133*).

The event horizon is defined as the boundary between points in the spacetime from which null geodesics can escape to infinity and points from which they cannot. The event horizon is therefore by definition a concept that depends on the entire spacetime. In the context of numerical simulations, this implies that an event horizon can only be computed if information about the entire spacetime is stored which results in large data sets even by contemporary standards. Nevertheless, event horizon finders have been developed in Refs. [278, 223]. For many purposes, however, it is more convenient to determine the existence of a horizon using data from a spatial hypersurface Σt only. Such a tool is available in the form of an AH. AHs are one of the most important diagnostic tools in NR and are reviewed in detail in the Living Reviews article [749]. It can be shown under the assumption of cosmic censorship and reasonable energy conditions, that the existence of an AH implies an event horizon whose cross section with Σt either lies outside the AH or coincides with it; see [406, 766] for details and proofs.

The key concept underlying the AH is that of a trapped surface defined as a surface where the expansion Θ ≡ ∇ μkμ of a congruence of outgoing null geodesics with tangent vector kμ satisfies Θ ≤ 0. A marginally trapped surface is defined as a surface where Θ = 0 and an AH is defined as the outermost marginally trapped surface on a spatial hypersurface Σt. Translated into the ADM variables, the condition Θ = 0 can be shown to lead to an elliptic equation for the unit normal direction sI to the D − 2-dimensional horizon surface

MN M N q DM sN − K + KMN s s = 0. (141 )
Here, qMN denotes the (D − 2 )-dimensional metric induced on the horizon surface. Numerical algorithms to solve this equation have been developed by several authors [374, 22, 747, 682, 748].

In the case of a static, spherically symmetric BH, it is possible to use the formula D− 2 Ahor = ΩD −2R S for the area of a D − 2 sphere to eliminate RS in Eq. (133*). We thus obtain an expression that relates the horizon area to a mass commonly referred to as the irreducible mass

D−3 (D − 2)ΩD −2 ( Ahor ) D−2 Mirr = ------------- ------ . (142 ) 16πG ΩD − 2
It is possible to derive the same expression in the more general case of a stationary BH, such as the Kerr BH in D = 4, or the Myers–Perry BH in D > 4.

The irreducible mass, as defined by Eq. 142*, is identical to the ADM mass for a static BH. This equation can be used to define the irreducible mass for stationary BHs as well [217*]. In D = 4 dimensions this becomes 16 πGM 2irr = Ahor. Furthermore, a rotating BH in D = 4 is described by a single spin parameter S and the BH mass consisting of rest mass and rotational energy has been shown by Christodoulou [217] to be given by

2 2 ---S2--- M = M irr + 4G2M 2. (143 ) irr
By adding the square of the linear momentum P 2 to the right-hand side of this equation we obtain the total energy of a spacetime containing a single BH with spin S and linear momentum P. In D = 4, Christodoulou’s formula (143*) can be used to calculate the spin from the equatorial circumference Ce and the horizon area according to [720*]
∘ ------ 2πAhor-= 1 + 1 − j2, (144 ) C2e
where j = S∕(GM 2) is the dimensionless spin parameter of the BH. Even though this relation is strictly valid only for the case of single stationary BHs, it provides a useful approximation in binary spacetimes as long as the BHs are sufficiently far apart.

It is a remarkable feature of BHs that their local properties such as mass and angular momentum can be determined in the way summarized here. In general it is not possible to assign in such a well-defined manner a local energy or momentum content to compact subsets of spacetimes due to the nonlinear nature of GR. For BHs, however, it is possible to derive expressions analogous to the ADM integrals discussed above, but now applied to the apparent horizon. Ultimately, this feature rests on the dynamic and isolated horizon framework; for more details see [281, 52] and the Living Reviews article by Ashtekar & Krishnan [53].

6.7.2 Gravitational-wave extraction

Probably the most important physical quantity to be extracted from dynamical BH spacetimes is the gravitational radiation. It is commonly extracted from numerical simulations in the form of either the Newman–Penrose scalar or a master function obtained through BH perturbation theory (see Section 5.2.1). Simulations using a characteristic formulation also facilitate wave extraction in the form of the Bondi mass loss formula. The Landau–Lifshitz pseudo-tensor [500], which has been generalized to D > 4 in [820*], has been used for gravitational radiation extraction in Ref. [700*] for studies of BH stability in higher dimensions; for applications in D = 4 see, e.g., [529]. Here, we will focus on the former two methods; wave extraction using the Bondi formalism is discussed in detail in Ref. [788].

Newman–Penrose scalar:
The formalism to extract GWs in the form of the Newman–Penrose scalar is currently fully understood only in D = 4 dimensions. Extension of this method is likely to require an improved understanding of the Goldberg–Sachs theorem in D > 4 which is subject to ongoing research [591]. The following discussion is therefore limited to D = 4 and we shall further focus on the case of asymptotically flat spacetimes. The Newman–Penrose formalism [575*] (see Section 5.2.1) is based on a tetrad of null vectors, two of them real and referred to as ℓ, k in this work, and two complex conjugate vectors referred to as m and ¯m; cf. Eq. (20*) and the surrounding discussion. Under certain conditions the projections of the Weyl tensor onto these tetrad directions may allow for a particularly convenient way to identify the physical properties of the spacetime. More specifically, the 10 independent components of the Weyl tensor are rearranged in the form of 5 complex scalars defined as (see, e.g., [573])

α β γ δ Ψ0 = − Cα βγδk m k m , Ψ1 = − Cα βγδk αℓβkγm δ, α β γ δ Ψ2 = − Cα βγδk m ¯m ℓ , Ψ3 = − Cα βγδk αℓβ ¯m γℓδ, α β γ δ Ψ4 = − Cα βγδℓ m¯ ℓ m¯ . (145 )
The identification of these projections with gravitational radiation is based on the work of Bondi et al. and Sachs [118, 667*] and the geometrical construction of Penrose [607] but crucially relies on a correct choice of the null tetrad in Eq. (145*) which needs to correspond to a Bondi frame. One example of this type, frequently considered in numerical applications, is the Kinnersley tetrad [469, 744]. More specifically, one employs a tetrad that converges to the Kinnersley tetrad as the spacetime approaches Petrov type D.16 Tetrads with this property are often referred to as quasi-Kinnersley tetrads and belong to a class of tetrads which are related to each other by spin/boost transformations; see [82, 572, 832*] and references therein. A particularly convenient choice consists in the transverse frame where Ψ = Ψ = 0 1 3 and the remaining scalars encode the ingoing gravitational radiation (Ψ0), the outgoing radiation (Ψ4) and the static or Coulomb part of the gravitational field (Ψ2). The construction of suitable tetrads in dynamic, numerically generated spacetimes represents a non-trivial task and is the subject of ongoing research (see, for example, [158, 510*, 571, 832]). For reasons already discussed in Section 6.6, extraction of gravitational waves is often performed at finite distance from the sources; but see Refs. [642, 60] for Cauchy-characteristic extraction that facilitates GW calculation at future null infinity. GW extraction at finite distances requires further ingredients which are discussed in more detail in [510]. These include a specific asymptotic behaviour of the Riemann tensor, the so-called peeling property [666, 667, 575], that outgoing null hypersurfaces define sequences of 2 S spheres which are conformal to unit spheres and a choice of coordinates that ensures appropriate fall-off of the metric components in the extraction frame.

Extraction of GWs at finite extraction radii rex is therefore affected by various potential errors. An attempt to estimate the uncertainty arising from the use of finite rex consists in measuring the GW signal at different values of the radius and analyzing its behaviour as the distance is increased. Convergence of the signal as 1∕r → 0 ex may then provide some estimate for the error incurred and improved results may be obtained through extrapolation to infinite rex; see, e.g., [124, 429*]. While such methods appear to work relatively well in practice (applying balance arguments together with measurements of BH horizon masses and the ADM mass or comparison with alternative extraction methods provide useful checks), it is important to bear in mind the possibility of systematic errors arising in the extraction of GWs using this method.

In the following discussion we will assume that the above requirements are met and describe a frequently used recipe that leads from the metric components of a numerical simulation to estimates of the energy and momenta contained in the gravitational radiation. The first step in the calculation of Ψ4 from the ADM metric is to construct the null tetrad. An approximation to a quasi-Kinnersley tetrad is given in terms of the unit timelike normal vector α n introduced in Section 6.1, and a triad i u, i v, i w of spatial vectors on each surface Σt constructed through Gram–Schmidt orthonormalization starting with

i i 2 2 i i m n u = [x, y,z], v = [xz,yz, − x − y ], w = 𝜖mnu v . (146 )
Here, 𝜖imn represents the three-dimensional Levi-Civita tensor on Σt and x,y,z are standard Cartesian coordinates. An orthonormal tetrad is then obtained from
α -1-- α α α -1-- α α α -1-- α α k = √2-(n + u ), ℓ = √2-(n − u ), m = √2--(v + iw ), (147 )
where time components of the spatial triad vectors vanish by construction.

Then, the calculation of Ψ4 from the ADM variables can be achieved either by constructing the spacetime metric from the spatial metric, lapse and shift vector and computing the spacetime Riemann or Weyl tensor through their definitions (see the preamble on “notation and conventions”). Alternatively, we can use the electric and magnetic parts of the Weyl tensor given by [334*]

Eαβ = ⊥ μα⊥ νβC μρνσnρnσ, Bα β = ⊥μα ⊥νβ∗C μρνσnρnσ, (148 )
where the ∗ denotes the Hodge dual. By using the Gauss–Codazzi equations (47*), one can express the electric and magnetic parts in vacuum in terms of the ADM variables according to17
mn kmn Eij = ℛij − γ (KijKmn − KimKjn ), Bij = γik𝜖 DmKnj. (149 )
In vacuum, the Weyl tensor is then given in terms of electric and magnetic parts by Eq. (3.10) in Ref. [334]. Inserting this relation together with (147*) and (149*) into the definition (145*) gives us the final expression for Ψ4 in terms of spatial variables
1 Ψ4 = − -[Emn (vmvn − wmwn ) − Bmn (vmwn + wmvn )] 2 + i[E (vmwn − wmvn ) + B (wmwn + vmvn )]. (150 ) 2 mn mn
The GW signal is often presented in the form of multipolar components ψ ℓm defined by projection of Ψ4 onto spherical harmonics of spin weight − 2  [356]
∑ ∫ ------ Ψ (t,𝜃,ϕ) = ψ (t)Y (− 2)(𝜃,ϕ ) ⇔ ψ (t) = Ψ (t,𝜃,ϕ)Y (− 2)(𝜃,ϕ) dΩ , (151 ) 4 lm lm lm 4 lm 2 lm
where the bar denotes the complex conjugate. The ψlm are often written in terms of amplitude and phase
iϕlm ψlm = Alme . (152 )
The amount of energy, linear and angular momentum carried by the GWs can be calculated from Ψ4 according to [663]
[ ∫ |∫ t |2 ] -dE- -r2- || &tidle;|| dt = lri→m∞ 16π Ω | −∞ Ψ4 dt| dΩ , (153 ) 2 [ 2 ∫ ||∫ t ||2 ] dPi-= − lim -r-- ℓi| Ψ4 dt&tidle;| dΩ , (154 ) dt r→∞ 16π Ω2 | −∞ | { [ ( ) ]} dJ r2 ∫ ( ∫ t ) ∫ t ∫ ˆt -- ---i= − lim ----Re Jˆi Ψ4 d&tidle;t Ψ4 d&tidle;t dˆt dΩ , dt r→∞ 16 π Ω2 −∞ −∞ −∞ (155 )
where
ℓi = [− sin 𝜃cos ϕ, − sin 𝜃sinϕ, − cos 𝜃], (156 ) ( 2i ) ˆJx = − sinϕ ∂𝜃 − cosϕ cot 𝜃∂ϕ − ----- , (157 ) ( sin)𝜃 2i ˆJy = cosϕ ∂𝜃 − sin ϕ cot 𝜃∂ϕ − ----- , (158 ) sin 𝜃 ˆJz = ∂ϕ. (159 )
In practice, one often starts the integration at the start of the numerical simulation (or shortly thereafter to avoid contamination from spurious GWs contained in the initial data) rather than at − ∞.

We finally note that the GW strain commonly used in GW data analysis is obtained from Ψ4 by integrating twice in time

∫ ( ∫ ) t &tidle;t h ≡ h+ − ih× = Ψ4 dˆt d&tidle;t. (160 ) −∞ −∞
h is often decomposed into multipoles in analogy to Eq. (151*). As before, the practical integration is often started at finite value rather than at − ∞. It has been noted that this process of integrating Ψ 4 twice in time is susceptible to large nonlinear drifts. These are due to fundamental difficulties that arise in the integration of finite-length, discretly sampled, noisy data streams which can be cured or at least mitigated by performing the integration in the Fourier instead of the time domain [644, 429*].

Perturbative wave extraction:
The basis of this approach to extract GWs from numerical simulations in D = 4 is the Regge–Wheeler–Zerilli–Moncrief formalism developed for the study of perturbations of spherically symmetric BHs. The assumption for applying this formalism to numerically generated spacetimes is that at sufficiently large distances from the GW sources, the spacetime is well approximated by a spherically symmetric background (typically Schwarzschild or Minkowski spacetime) plus non-spherical perturbations. These perturbations naturally divide into odd and even multipoles which obey the Regge–Wheeler [641] (odd) and the Zerilli [830] (even) equations respectively (see Section 5.2.1). Moncrief [555] developed a gauge-invariant formulation for these perturbations in terms of a master function which obeys a wave-type equation with a background dependent scattering potential; for a review and applications of this formalism see for example [566, 721, 643]. An extension of this formalism to higher-dimensional spacetimes has been developed by Kodama & Ishibashi [479], and is discussed in Section 5.2.3. This approach has been used to develop wave extraction from NR simulations in D > 4 with SO (D − 2) symmetry [797*]. In particular, it has been applied to the extraction of GWs from head-on collisions of BHs. As in our discussion of formulations of the Einstein equations in higher dimensions in Section 6.2, it turns out useful to introduce coordinates that are adapted to the rotational symmetry on a D− 2 S sphere. Here, we choose spherical coordinates for this purpose which we denote by a (t,r,𝜗,𝜃,ϕ ) where a = 4, ...,D − 1; we use the same convention for indices as in Section 6.2.

We then assume that in the far-field region, the spacetime is perturbatively close to a spherically symmetric BH background given in D dimensions by the Tangherlini [740] metric

2 2 2 −1 2 2 [ 2 2 2 2 ] ds(0) = − A(r) dt + A (r) dr + r d𝜗 + sin 𝜗 (d𝜃 + sin 𝜃d ΩD −4) , (161 )
where
D −3 A (r) = 1 − RS---, (162 ) rD− 3
and the Schwarzschild radius RS is related to the BH mass through Eq. (133*). For a spacetime with SO (D − 3) isometry the perturbations away from the background (161*) are given by
ds2(1) = hAB dxA dxB + hA 𝜗dxA d𝜗 + h 𝜗𝜗d𝜗2 + h 𝜃𝜃 dΩD −3, (163 )
where we introduce early upper case Latin indices A, B, ...= 0,1 and A x = (t,r). The class of axisymmetric spacetimes considered in [797*] obeys SO (D − 2) isometry which can be shown to imply that hA𝜃 = h 𝜗𝜃 = 0 and that the remaining components of h in Eq. (163*) only depend on the coordinates (t,r,𝜗). As a consequence, only the perturbations which we have called in Section 5.2.3 “scalar” are non-vanishing, and are expanded in tensor spherical harmonics; cf. Section II C in Ref. [797*].

As discussed in Section 5.2.3, the metric perturbations, decomposed in tensor harmonics, can be combined in a gauge-invariant master function Φ ℓm. From the master function, we can calculate the GW energy flux and the total radiated energy as discussed in Section 5.2.3.

6.7.3 Diagnostics in asymptotically AdS spacetimes

The gauge/gravity duality, or AdS/CFT correspondence (see Section 3.3.1), relates gravity in asymptotically AdS spacetimes to conformal field theories on the boundary of this spacetime. A key ingredient of the correspondence is the relation between fields interacting gravitationally in the bulk spacetime and expectation values of the field theory on the boundary. Here we restrict our attention to the extraction of the expectation values of the energy-momentum tensor ⟨TIJ⟩ of the field theory from the fall-off behaviour of the AdS metric.

Through the AdS/CFT correspondence, the expectation values ⟨TIJ⟩ of the field theory are given by the quasi-local Brown–York [139] stress-energy tensor and thus are directly related to the bulk metric. Following [253*], it is convenient to consider the (asymptotically AdS) bulk metric in Fefferman–Graham [314] coordinates

2 μ ν L2-[ 2 I J] ds = gμν dx dx = r2 dr + γIJ dx dx , (164 )
where
γ = γ (r,xI) = γ + r2γ + ...+ rdγ + h rd log r2 + 𝒪(rd+1). (165 ) IJ IJ (0)IJ (2)IJ (d)IJ (d)IJ
Here d ≡ D − 1, the γ(a)IJ and h(d)IJ are functions of the boundary coordinates xi, the logarithmic term only appears for even d and powers of r are exclusively even up to order d − 1. As shown in Ref. [253*], the vacuum expectation value of the CFT momentum tensor for d = 4 is then obtained from
3 { 4L-- 1- [ 2 KM LN ] ⟨TIJ⟩ = 16 π γ (4)IJ − 8 γ(0)IJ γ(2) − γ(0) γ(0)γ(2)KLγ(2)MN } − 1-γ(2)IMγ (2)JM + 1γ (2)IJγ(2) , (166 ) 2 4
and γ(2)IJ is determined in terms of γ(0)IJ. The dynamical freedom of the CFT is thus encapsulated in the fourth-order term γ(4)IJ. If γ(0)IJ = ηIJ, for r → 0 the metric (164*) asymptotes to the AdS metric in Poincaré coordinates (131*).

The Brown–York stress tensor is also the starting point for an alternative method to extract the ⟨TIJ⟩ that does not rely on Fefferman-Graham coordinates. It is given by

2 δS T μν = √-------grav-, (167 ) − γ δ γμν
where we have foliated the D-dimensional spacetime into timelike hypersurfaces Σr in analogy to the foliation in terms of spacelike hypersurfaces Σt in Section 6.1.1. The spacetime metric is given by
ds2 = α2 dr2 + γ (dxI + βI dr )(dxJ + βJ dr). (168 ) IJ
In analogy to the second fundamental form K αβ in Section 6.1.1, we define the extrinsic curvature on Σr by
Θ μν ≡ − 1(∇ μnν + ∇ νnμ), (169 ) 2
where nμ denotes the outward pointing normal vector to Σ r. Reference [67*] provides a method to cure divergencies that appear in the Brown–York tensor when the boundary is pushed to infinity by adding counterterms to the action Sgrav. This work discusses asymptotically AdS spacetimes of different dimensions. For AdS5, the procedure results in
[ ] μν -1- μν μν 3- μν L- μν T = 8π Θ − Θ γ − L γ − 2 𝒢 , (170 )
where 𝒢μν = ℛ μν − ℛ γμν∕2 is the Einstein tensor associated with the induced metric γμν. Applied to the AdS5 metric in global coordinates, this expression gives a non-zero energy-momentum tensor Tμν which, translated into the expectation values μν ⟨T ⟩, can be interpreted as the Casimir energy of a quantum field theory on the spacetime with topology ℝ × S3 [67]. This Casimir energy is non-dynamical and in numerical applications to the AdS/CFT correspondence may simply be subtracted from Tμν; see, e.g., [70*].

The role of additional (e.g., scalar) fields in the AdS/CFT dictionary is discussed, for example, in Refs. [253, 705].


  Go to previous page Scroll to top Go to next page