1 Introduction
During the last 25 years one of the greatest achievements in classical general relativity is certainly the proof of the positivity of the total gravitational energy, both at spatial and null infinity. It is precisely its positivity that makes this notion not only important (because of its theoretical significance), but a useful tool as well in the everyday practice of working relativists. This success inspired the more ambitious claim to associate energy (or rather energy-momentum and, ultimately, angular momentum too) to extended but finite spacetime domains, i. e. at the quasi-local level. Obviously, the quasi-local quantities could provide a more detailed characterization of the states of the gravitational ‘field’ than the global ones, so they (together with more general quasi-local observables) would be interesting in their own right. Moreover, finding an appropriate notion of energy-momentum and angular momentum would be important from the point of view of applications as well. For example, they may play a central role in the proof of the full Penrose inequality (as they have already played in the proof of the Riemannian version of this inequality). The correct, ultimate formulation of black hole thermodynamics should probably be based on quasi-locally defined internal energy, entropy, angular momentum etc. In numerical calculations conserved quantities (or at least those for which balance equations can be derived) are used to control the errors. However, in such calculations all the domains are finite, i. e. quasi-local. Therefore, a solid theoretical foundation of the quasi-local conserved quantities is needed. However, contrary to the high expectations of the eighties, finding an appropriate quasi-local notion of energy-momentum has proven to be surprisingly difficult. Nowadays, the state of the art is typically postmodern: Although there are several promising and useful suggestions, we have not only no ultimate, generally accepted expression for the energy-momentum and especially for the angular momentum, but there is no consensus in the relativity community even on general questions (for example, what should we mean e. g. by energy-momentum: Only a general expression containing arbitrary functions, or rather a definite one free of any ambiguities, even of additive constants), or on the list of the criteria of reasonableness of such expressions. The various suggestions are based on different philosophies, approaches and give different results in the same situation. Apparently, the ideas and successes of one construction have only very little influence on other constructions. The aim of the present paper is therefore twofold. First, to collect and review the various specific suggestions, and, second, to stimulate the interaction between the different approaches by clarifying the general, potentially common points, issues, questions. Thus we wanted to write not only a ‘who-did-what’ review, but primarily we would like to concentrate on the understanding of the basic questions (such as why should the gravitational energy-momentum and angular momentum, or, more generally, any observable of the gravitational ‘field’, be necessarily quasi-local) and ideas behind the various specific constructions. Consequently, one-third of the present review is devoted to these general questions. We review the specific constructions and their properties only in the second part, and in the third part we discuss very briefly some (potential) applications of the quasi-local quantities. Although this paper is basically a review of known and published results, we believe that it contains several new elements, observations, suggestions etc. Surprisingly enough, most of the ideas and concepts that appear in connection with the gravitational energy-momentum and angular momentum can be introduced in (and hence can be understood from) the theory of matter fields in Minkowski spacetime. Thus, in subsection 2.1 , we review the Belinfante–Rosenfeld procedure that we will apply to gravity in section 3 , introduce the notion of quasi-local energy-momentum and angular momentum of the matter fields and discuss their properties.The philosophy of quasi-locality in general relativity will be demonstrated in Minkowski spacetime where the energy-momentum and angular momentum of the matter fields are treated quasi-locally. Then we turn to the difficulties of gravitational energy-momentum and angular momentum, and we clarify why the gravitational observables should necessarily be quasi-local. The tools needed to construct and analyze the quasi-local quantities are reviewed in the fourth section. This closes the first, the general part of the review. The second part is devoted to the discussion of the specific constructions (sections 5 – 12 ). Since most of the suggestions are constructions, they cannot be given as a short mathematical definition. Moreover, there are important physical ideas behind them, without which the constructions may appear ad hoc. Thus we always try to explain these physical pictures, the motivations and interpretations. Although the present paper is intended to be a non-technical review, the explicit mathematical definitions of the various specific constructions will always be given. Then the properties and the applications are usually summarized only in a nutshell. Sometimes we give a review on technical aspects too, without which it would be difficult to understand even some of the conceptual issues. The list of references connected with this second part is intended to be complete. We apologize to all those whose results were accidentally left out. The list of the (actual and potential) applications of the quasi-local quantities, discussed in section 13 , is far from being complete, and might be a little bit subjective. Here we consider the calculation of gravitational energy transfer, applications in black hole physics and a quasi-local characterization of the pp-wave metrics. We close this paper with a discussion of the successes and deficiencies of the general and (potentially) viable constructions. In contrast to the positivistic style of sections 5 – 12 , section 14 (as well as the choice for the matter of sections 2 – 4 ) reflects our own personal interest and view of the subject. The theory of quasi-local observables in general relativity is far from being complete. The most important open problem is still the trivial one: ‘Find quasi-local energy-momentum and angular momentum expressions satisfying the points of the lists of subsection 4.3 ’. Several specific open questions in connection with the specific definitions are raised both in the corresponding sections and in section 14 , which could be worked out even by graduate students. On the other hand, any of their application to solve physical/geometrical problems (e. g. to some mentioned in section 13 ) would be a real success. In the present paper we adopt the abstract index formalism. The signature of the spacetime metric is ; and the curvature and Ricci tensors and the curvature scalar of the covariant derivative are defined by , and , respectively. Hence Einstein’s equations take the form , where is Newton’s gravitational constant (and the speed of light is ). However, in subsections 13.3 and 13.4 we use the traditional system.2 Energy-Momentum and Angular Momentum of Matter Fields
2.1 Energy-momentum and angular momentum density of matter fields
2.1.1 The symmetric energy-momentum tensor
It is a widely accepted view (appearing e. g. in excellent, standard textbooks on general relativity, too) that the canonical energy-momentum and spin tensors are well defined and have relevance only in flat spacetime, and hence usually are underestimated and abandoned. However, it is only the analog of these canonical quantities that can be associated with gravity itself. Thus first we introduce these quantities for the matter fields in a general curved spacetime. To specify the state of the matter fields operationally two kinds of devices are needed: The first measures the value of the fields, while the other measures the spatio-temporal location of the first. Correspondingly, the fields on the manifold of events can be grouped into two sharply distinguished classes: The first contains the matter field variables, e. g. finitely many type tensor fields ; whilst the other contains the fields specifying the spacetime geometry, i. e. the metric in Einstein’s theory. Suppose that the dynamics of the matter fields is governed by Hamilton’s principle specified by a Lagrangian : If is the volume integral of on some open domain with compact closure then the equations of motion are , the Euler–Lagrange equations. The symmetric (or dynamical) energy-momentum tensor is defined (and is given explicitly) by(1) |
(2) |
2.1.2 The canonical Noether current
Suppose that the Lagrangian is weakly diffeomorphism invariant in the sense that for any vector field and the corresponding local 1-parameter family of diffeomorphisms one has for some one parameter family of vector fields . ( is called diffeomorphism invariant if , e. g. when is a scalar.) Let be any smooth vector field on . Then, calculating the divergence to determine the rate of change of the action functional along the integral curves of , by a tedious but straightforward computation one can derive the so-called Noether identity: , where denotes the Lie derivative along and , the so-called Noether current, is given explicitly by(3) |
(4) |
2.2 Quasi-local energy-momentum and angular momentum of the matter fields
In the next section we will see that well defined (i. e. gauge invariant) energy-momentum and angular momentum density cannot be associated with the gravitational ‘field’, and if we want to talk not only about global gravitational energy-momentum and angular momentum, then these quantities must be assigned to extended but finite spacetime domains. In the light of modern quantum field theoretical investigations it has become clear that all physical observables should be associated with extended but finite spacetime domains [164, 163] . Thus observables are always associated with open subsets of spacetime whose closure is compact, i. e. they are quasi-local. Quantities associated with spacetime points or with the whole spacetime are not observable in this sense. In particular, global quantities, such as the total energy or electric charge, should be considered as the limit of quasi-locally defined quantities. Thus the idea of quasi-locality is not new in physics. Although apparently in classical non-gravitational physics this is not obligatory, we adopt this view in talking about energy-momentum and angular momentum even of classical matter fields in Minkowski spacetime. Originally the introduction of these quasi-local quantities was motivated by the analogous gravitational quasi-local quantities [338, 342] . Since, however, many of the basic concepts and ideas behind the various gravitational quasi-local energy-momentum and angular momentum definitions can be understood from the analogous non-gravitational quantities in Minkowski spacetime, we devote the present subsection to the discussion of them and their properties.2.2.1 The definition of the quasi-local quantities
To define the quasi-local conserved quantities in Minkowski spacetime, first observe that for any Killing vector the 3-form is closed, and hence, by the triviality of the third de Rham cohomology class, , it is exact: For some 2-form we have . may be called a superpotential for the conserved current 3-form . (The existence of globally defined superpotentials can be proven even without using the Poincare lemma [370] .) If is (the dual of) another superpotential for the same current , then by and the dual superpotential is unique up to the addition of an exact 2-form. If therefore is any closed orientable spacelike 2-surface in the Minkowski spacetime then the integral of on is free from this ambiguity. Thus if is any smooth compact spacelike hypersurface with smooth 2-boundary , then(5) |
(6) |
(7) |
2.2.2 Hamiltonian introduction of the quasi-local quantities
In the standard Hamiltonian formulation of the dynamics of the classical matter fields on a given (not necessarily flat) spacetime (see for example [206, 377] and references therein) the configuration and momentum variables, and , respectively, are fields on a connected 3-manifold , which is interpreted as the typical leaf of a foliation of the spacetime. The foliation can be characterized on by a function , called the lapse. The evolution of the states in the spacetime is described with respect to a vector field (‘evolution vector field’ or ’general time axis’), where is the future directed unit normal to the leaves of the foliation and is some vector field, called the shift, being tangent to the leaves. If the matter fields have gauge freedom, then the dynamics of the system is constrained: Physical states can be only those that are on the constraint surface, specified by the vanishing of certain functions , , of the canonical variables and their derivatives up to some finite order, where is the covariant derivative operator in . Then the time evolution of the states in the phase space is governed by the Hamiltonian, which has the form(8) |
Sometimes in the literature this requirement is introduced as some new principle in the Hamiltonian formulation of the fields, but its real content is not more than to ensure that the Hamilton equations coincide with the field equations.
2.2.3 Properties of the quasi-local quantities
Suppose that the matter fields satisfy the dominant energy condition. Then is also non-negative for any non-spacelike , and, obviously, is zero precisely when on , and hence, by the conservation laws (see for example page 94 of [170] ), on the whole domain of dependence . Obviously, if and only if is null on . Then by the dominant energy condition it is a future pointing vector field on , and holds. Therefore, on has a null eigenvector with zero eigenvalue, i. e. its algebraic type on is pure radiation. The properties of the quasi-local quantities based on in Minkowski spacetime are, however, more interesting. Namely, assuming that the dominant energy condition is satisfied, one can prove [338, 342] that2.2.4 Global energy-momenta and angular momenta
If extends either to spacelike or future null infinity, then, as is well known, the existence of the limit of the quasi-local energy-momentum can be ensured by slightly faster than (for example by ) fall-off of the energy-momentum tensor, where is any spatial radial distance. However, the finiteness of the angular momentum and centre-of-mass is not ensured by the fall-off. Since the typical fall-off of – for example for the electromagnetic field – is , we may not impose faster than this, because otherwise we would exclude the electromagnetic field from our investigations. Thus, in addition to the fall-off, six global integral conditions for the leading terms of must be imposed. At the spatial infinity these integral conditions can be ensured by explicit parity conditions, and one can show that the ‘conservation equations’ (as evolution equations for the energy density and momentum density) preserve these fall-off and parity conditions [348] . Although quasi-locally the vanishing of the mass does not imply the vanishing of the matter fields themselves (the matter fields must be pure radiative field configurations with plane wave fronts), the vanishing of the total mass alone does imply the vanishing of the fields. In fact, by the vanishing of the mass the fields must be plane waves, furthermore by they must be asymptotically vanishing at the same time. However, a plane wave configuration can be asymptotically vanishing only if it is vanishing.2.2.5 Quasi-local radiative modes of the matter fields
By the results of the previous subsection the vanishing of the quasi-local mass, associated with a closed spacelike 2-surface , implies that the matter must be pure radiation on a four dimensional globally hyperbolic domain . Thus characterizes ‘simple’, ‘elementary’ states of the matter fields. In the present subsection we raise the question if these states on can be characterized completely by data on the 2-surface or not. For the (real or complex) linear massless scalar field and the Maxwell field, represented by the symmetric spinor field , the condition implies that they are completely determined on the whole by the value of these fields on (and is necessarily null: , where is a complex function and is a constant spinor field such that ). Similarly, the null linear zero-rest-mass fields on with any spin, , are completely determined by their value on (assuming, for the sake of simplicity, that is future and past convex in the sense of subsection 4.1.3 below). Technically, these results are based on the unique complex analytic structure of the 2-surfaces foliating , where ; and by the field equations the complex functions turn out to be holomorphic. Thus they are completely determined on (and hence on too) by their value on [342] . Since the independent radiative modes of the linear zero-rest-mass fields are just the -type solutions of their field equations, characterized by the null wavevector and the amplitude in their Fourier expansion, their radiative modes can equivalently be characterized by not only the familiar initial data on three dimensional spacelike hypersurfaces, but by data on closed 2-surfaces as well. Intuitively this result should be clear, because the two radiative (transversal) degrees of freedom of each Fourier mode may be expected to be characterized by two real functions on a two dimensional ‘screen’. This fact may be a specific manifestation in the classical non-gravitational physics of the holographic principle (see subsection 13.4.2 ). A more detailed discussion of these issues will be given elsewhere.3 On the Energy-Momentum and Angular Momentum of Gravitating Systems
3.1 On the gravitational energy-momentum and angular momentum density: the difficulties
3.1.1 The root of the difficulties
The action for the matter fields was a functional of both kinds of fields, thus one could take the variational derivatives both with respect to and . The former gave the field equations, while the latter defined the symmetric energy-momentum tensor. Moreover, provided a metrical geometric background, in particular a covariant derivative, for carrying out the analysis of the matter fields. The gravitational action is, on the other hand, a functional of the metric alone, and its variational derivative with respect to yields the gravitational field equations. The lack of any further geometric background for describing the dynamics of can be traced back to the principle of equivalence [21] , and introduces a huge gauge freedom in the dynamics of because that should be formulated on a bare manifold: The physical spacetime is not simply a manifold endowed with a Lorentzian metric , but the isomorphism class of such pairs, where and are considered to be equivalent for any diffeomorphism of onto itself. Thus we do not have, even in principle, any gravitational analog of the symmetric energy-momentum tensor of the matter fields. In fact, by its very definition, is the source-current for gravity, like the current in Yang–Mills theories (defined by the variational derivative of the action functional of the particles, e. g. of the fermions, interacting with a Yang–Mills field ), rather than energy-momentum. The latter is represented by the Noether currents associated with special spacetime displacements. Thus, in spite of the intimate relation between and the Noether currents, the proper interpretation of is only the source density for gravity, and hence it is not the symmetric energy-momentum tensor whose gravitational counterpart must be searched for. In particular, the Bel–Robinson tensor , given in terms of the Weyl spinor, (and its generalizations introduced by Senovilla [318, 317] ), being a quadratic expression of the curvature (and its derivatives), is (are) expected to represent only ‘higher order’ gravitational energy-momentum. In fact, the physical dimension of the Bel–Robinson ‘energy-density’ is , and hence (in the traditional units) there are no powers and such that would have energy-density dimension. Here is the speed of light and is Newton’s gravitational constant. As we will see, the Bel–Robinson ‘energy-momentum density’ appears naturally in connection with the quasi-local energy-momentum and spin-angular momentum expressions for small spheres only in higher order terms. Therefore, if we want to associate energy-momentum and angular momentum with the gravity itself in a Lagrangian framework, then it is the gravitational counterpart of the canonical energy-momentum and spin tensors and the canonical Noether current built from them that should be introduced. Hence it seems natural to apply the Lagrange–Belinfante–Rosenfeld procedure, sketched in the previous section, to gravity too [55, 56, 308, 188, 189, 336] .Since we do not have a third kind of device to specify the spatio-temporal location of the devices measuring the spacetime geometry, we do not have any further operationally defined, maybe non-dynamical background, just in accordance with the principle of equivalence. If there were some non-dynamical background metric on , then by requiring we could reduce the almost arbitrary diffeomorphism (essentially four arbitrary functions of four variables) to a transformation depending on at most ten parameters.
3.1.2 Pseudotensors
The lack of any background geometric structure in the gravitational action yields, first, that any vector field generates a symmetry of the matter plus gravity system. Its second consequence is the need for an auxiliary derivative operator, e. g. the Levi-Civita covariant derivative coming from an auxiliary, non-dynamical background metric (see for example [225, 302] ), or a background (usually torsion free, but not necessarily flat) connection (see for example [209] ), or the partial derivative coming from a local coordinate system (see for example [365] ). Though the natural expectation would be that the final results be independent of these background structures, as is well known, the results do depend on them. In particular [336] , for Hilbert’s second order Lagrangian in a fixed local coordinate system and derivative operator instead of ( 4 ) gives precisely Møller’s energy-momentum pseudotensor , which was defined originally through the superpotential equation , where is the so-called Møller superpotential [260] . (For another simple and natural introduction of Møller’s energy-momentum pseudotensor see [101] .) For the spin pseudotensor ( 2 ) gives which is in fact only pseudotensorial. Similarly, the contravariant form of these pseudotensors and the corresponding canonical Noether current are also pseudotensorial. We saw in subsection 2.1.2 that a specific combination of the canonical energy-momentum and spin tensors gave the symmetric energy-momentum tensor, which is gauge invariant even if the matter fields have gauge freedom, and one might hope that the analogous combination of the energy-momentum and spin pseudotensors gives a reasonable tensorial energy-momentum density for the gravitational field. The analogous expression is, in fact, tensorial, but unfortunately it is just minus the Einstein tensor [336, 337] . Therefore, to use the pseudotensors a ‘natural’ choice for a ‘preferred’ coordinate system would be needed. This could be interpreted as a gauge choice, or reference configuration. A further difficulty is that the different pseudotensors may have different (potential) significance. For example, for any fixed Goldberg’s -th symmetric pseudotensor is defined by (which, for , reduces to the Landau–Lifshitz pseudotensor, the only symmetric pseudotensor which is a quadratic expression of the first derivatives of the metric) [158] . However, by Einstein’s equations this definition implies that . Hence what is (coordinate-)divergence-free (i. e. ‘pseudo-conserved’) cannot be interpreted as the sum of the gravitational and matter energy-momentum densities. Indeed, the latter is , while the second term in the divergence equation has an extra weight . Thus there is only one pseudotensor in this series, , which satisfies the ‘conservation law’ with the correct weight. On the other hand, the pseudotensors coming from some action (the ‘canonical pseudotensors’) appear to be free of this kind of difficulties (see also [336, 337] ). Classical excellent reviews on these (and several other) pseudotensors are [365, 57, 8, 159] , and for some recent ones (using background geometric structures) see for example [133, 134, 77, 150, 151, 222, 302] .) We return to the discussion of pseudotensors in subsections 3.3.1 and 11.3.3 .Since Einstein’s Lagrangian is only weakly diffeomorphism invariant, the situation would even be worse if we used Einstein’s Lagrangian. The corresponding canonical quantities would still be coordinate dependent, though in certain ‘natural’ coordinate system they yield reasonable results (see for example [2] and references therein).
3.1.3 Strategies to avoid pseudotensors I: Background metrics/connections
One way of avoiding the use of the pseudotensorial quantities is to introduce an explicit background connection [209] or background metric [307, 223, 227, 225, 224, 301, 131] . The advantage of this approach would be that we could use the background not only to derive the canonical energy-momentum and spin tensors, but to define the vector fields as the symmetry generators of the background. Then the resulting Noether currents are without doubt tensorial. However, they depend explicitly on the choice of the background connection or metric not only through : The canonical energy-momentum and spin tensors themselves are explicitly background-dependent. Thus, again, the resulting expressions would have to be supplemented by a ‘natural’ choice for the background, and the main question is how to find such a ‘natural’ reference configuration from the infinitely many possibilities.3.1.4 Strategies to avoid pseudotensors II: The tetrad formalism
In the tetrad formulation of general relativity the -orthonormal frame fields , , are chosen to be the gravitational field variables [368, 230] . Re-expressing the Hilbert Lagrangian (i. e. the curvature scalar) in terms of the tetrad field and its partial derivatives in some local coordinate system, one can calculate the canonical energy-momentum and spin by ( 4 ) and ( 2 ), respectively. Not surprisingly at all, we recover the pseudotensorial quantities that we obtained in the metric formulation above. However, as realized by Møller [261] , the use of the tetrad fields as the field variables instead of the metric makes it possible to introduce a first order, scalar Lagrangian for Einstein’s field equations: If , the Ricci rotation coefficients, then Møller’s tetrad Lagrangian is(9) |
(10) |
(11) |
3.1.5 Strategies to avoid pseudotensors III: Higher derivative currents
Giving up the paradigm that the Noether current should depend only on the vector field and its first derivative – i. e. if we allow a term to be present in the Noether current ( 3 ) even if the Lagrangian is diffeomorphism invariant – one naturally arrives at Komar’s tensorial superpotential and the corresponding Noether current ([235] , see also [57] ). Although its independence of any background structure (viz. its tensorial nature) and uniqueness property (see Komar [235] quoting Sachs) is especially attractive, the vector field is still to be determined.3.2 On the global energy-momentum and angular momentum of gravitating systems: the successes
As is well known, in spite of the difficulties with the notion of the gravitational energy-momentum density discussed above, reasonable total energy-momentum and angular momentum can be associated with the whole spacetime provided it is asymptotically flat. In the present subsection we recall the various forms of them. As we will see, most of the quasi-local constructions are simply ‘quasi-localizations’ of the total quantities. Obviously, the technique used in the ‘quasi-localization’ does depend on the actual form of the total quantities, yielding mathematically inequivalent definitions for the quasi-local quantities. We return to the discussion of the tools needed in the quasi-localization procedures in subsections 4.2 and 4.3 . Classical, excellent reviews of global energy-momentum and angular momentum are [147, 159, 14, 276, 375, 299] , and a recent review of conformal infinity (with special emphasis on its applicability in numerical relativity) is [140] . Reviews of the positive energy proofs from the first third of the eighties are [196, 300] .3.2.1 Spatial infinity: Energy-momentum
There are several mathematically inequivalent definitions of asymptotic flatness at spatial infinity [147, 328, 22, 47, 144] . The traditional definition is based on the existence of a certain asymptotically flat spacelike hypersurface. Here we adopt this definition, which is probably the weakest one in the sense that the spacetimes that are asymptotically flat in the sense of any reasonable definition are asymptotically flat in the traditional sense too. A spacelike hypersurface will be called -asymptotically flat if for some compact set the complement is diffeomorphic to minus a solid ball, and there exists a (negative definite) metric on , which is flat on , such that the components of the difference of the physical and the background metrics, , and of the extrinsic curvature in the -Cartesian coordinate system fall off as and , respectively, for some and [304, 46] . These conditions make it possible to introduce the notion of asymptotic spacetime Killing vectors, and to speak about asymptotic translations and asymptotic boost–rotations. together with the metric and extrinsic curvature is called the asymptotic end of . In a more general definition of asymptotic flatness is allowed to have finitely many such ends. As is well known, finite and well defined ADM energy-momentum [10, 12, 11, 13] can be associated with any -asymptotically flat spacelike hypersurface if by taking the value on the constraint surface of the Hamiltonian , given for example in [304, 46] , with the asymptotic translations (see [109, 36, 278, 110] ). In its standard form this is the limit of a 2-surface integral of the first derivatives of the induced 3-metric and of the extrinsic curvature for spheres of large coordinate radius . The ADM energy-momentum is an element of the space dual to the space of the asymptotic translations, and transforms as a Lorentzian 4-vector with respect to asymptotic Lorentz transformations of the asymptotic Cartesian coordinates. The traditional ADM approach to the introduction of the conserved quantities and the Hamiltonian analysis of general relativity is based on the 3+1 decomposition of the fields and the spacetime. Thus it is not a priori clear that the energy and spatial momentum form a Lorentz vector (and the spatial angular momentum and centre-of-mass, discussed below, form an anti-symmetric tensor). One had to check a posteriori that the conserved quantities obtained in the 3+1 form are, in fact, Lorentz-covariant. To obtain manifestly Lorentz-covariant quantities one should not do the 3+1 decomposition. Such a manifestly Lorentz-covariant Hamiltonian analysis was suggested first by Nester [269] , and he was able to recover the ADM energy-momentum in a natural way (see also subsection 11.3 below). Another form of the ADM energy-momentum is based on Møller’s tetrad superpotential [159] : Taking the flux integral of the current on the spacelike hypersurface , by ( 11 ) the flux can be rewritten as the limit of the 2-surface integral of Møller’s superpotential on spheres of large with the asymptotic translations . Choosing the tetrad field to be adapted to the spacelike hypersurface and assuming that the frame tends to a constant Cartesian one as , the integral reproduces the ADM energy-momentum. The same expression can be obtained by following the familiar Hamiltonian analysis using the tetrad variables too: By the standard scenario one can construct the basic Hamiltonian [271] . This Hamiltonian, evaluated on the constraints, turns out to be precisely the flux integral of on . A particularly interesting and useful expression for the ADM energy-momentum is possible if the tetrad field is considered to be a frame field built from a normalized spinor dyad , , on which is asymptotically constant (see subsection 4.2.3 below). (Thus underlined capital Roman indices are concrete name spinor indices.) Then, for the components of the ADM energy-momentum in the constant spinor basis at infinity, Møller’s expression yields the limit of(12) |
(13) |
(14) |
(15) |
3.2.2 Spatial infinity: Angular momentum
The value of the Hamiltonian of Beig and O Murchadha [46] together with the appropriately defined asymptotic rotation–boost Killing vectors [348] define the spatial angular momentum and centre-of-mass, provided and, in addition to the familiar fall-off conditions, certain global integral conditions are also satisfied. These integral conditions can be ensured by the explicit parity conditions of Regge and Teitelboim [304] on the leading nontrivial parts of the metric and extrinsic curvature : The components in the Cartesian coordinates of the former must be even and the components of latter must be odd parity functions of (see also [46] ). Thus in what follows we assume that . Then the value of the Beig–O Murchadha Hamiltonian parameterized by the asymptotic rotation Killing vectors is the spatial angular momentum of Regge and Teitelboim [304] , while that parameterized by the asymptotic boost Killing vectors deviate from the centre-of-mass of Beig and O Murchadha [46] by a term which is the spatial momentum times the coordinate time. (As Beig and O Murchadha pointed out [46] , the centre-of-mass of Regge and Teitelboim is not necessarily finite.) The spatial angular momentum and the new centre-of-mass form an anti-symmetric Lorentz 4-tensor, which transforms in the correct way under the 4-translation of the origin of the asymptotically Cartesian coordinate system, and it is conserved by the evolution equations [348] . The centre-of-mass of Beig and O Murchadha was reexpressed recently [41] as the limit of 2-surface integrals of the curvature in the form ( 15 ) with proportional to the lapse times , where is the induced 2-metric on (see subsection 4.1.1 below). A geometric notion of centre-of-mass was introduced by Huisken and Yau [203] . They foliate the asymptotically flat hypersurface by certain spheres with constant mean curvature. By showing the global uniqueness of this foliation asymptotically, the origin of the leaves of this foliation in some flat ambient Euclidean space defines the centre-of-mass (or rather ‘centre-of-gravity’) of Huisken and Yau. However, no statement on its properties is proven. In particular, it would be interesting to see whether or not this notion of centre-of-mass coincides, for example, with that of Beig and O Murchadha. The Ashtekar–Hansen definition for the angular momentum is introduced in their specific conformal model of the spatial infinity as a certain 2-surface integral near infinity. However, their angular momentum expression is finite and unambiguously defined only if the magnetic part of the spacetime curvature tensor (with respect to the timelike level hypersurfaces of the conformal factor) falls off faster than would follow from the fall-off of the metric (but they do not have to impose any global integral, e. g. a parity condition) [22, 14] . If the spacetime admits a Killing vector of axis-symmetry, then the usual interpretation of the corresponding Komar integral is the appropriate component of the angular momentum (see for example [369] ). However, the value of the Komar integral is twice the expected angular momentum. In particular, if the Komar integral is normalized such that for the Killing field of stationarity in the Kerr solution the integral is , for the Killing vector of axis-symmetry it is instead of the expected (‘factor-of-two anomaly’) [223] .3.2.3 Null infinity: Energy-momentum
The study of the gravitational radiation of isolated sources led Bondi to the observation that the 2-sphere integral of a certain expansion coefficient of the line element of a radiative spacetime in an asymptotically retarded spherical coordinate system behaves as the energy of the system at the retarded time : This notion of energy is not constant in time, but decreases with , showing that gravitational radiation carries away positive energy (‘Bondi’s mass-loss’) [69, 70] . The set of transformations leaving the asymptotic form of the metric invariant was identified as a group, nowadays known as the BMS group, having a structure very similar to that of the Poincare group [310] . The only difference is that while the Poincare group is a semidirect product of the Lorentz group and a four dimensional commutative group (of translations), the BMS group is the semidirect product of the Lorentz group and an infinite dimensional commutative group, called the group of the supertranslations. A four parameter subgroup in the latter can be identified in a natural way as the group of the translations. Just at the same time the study of asymptotic solutions of the field equations led Newman and Unti to another concept of energy at null infinity [277] . However, this energy (nowadays known as the Newman–Unti energy) does not seem to have the same significance as the Bondi (or Bondi–Sachs [299] or Trautman–Bondi [112, 113, 111] ) energy, because its monotonicity can be proven only between special, e. g. stationary, states. The Bondi energy, which is the time component of a Lorentz vector, the so-called Bondi–Sachs energy-momentum, has a remarkable uniqueness property [112, 113] . Without additional conditions on , Komar’s expression does not reproduce the Bondi–Sachs energy-momentum in non-stationary spacetimes either [376, 159] : For the ‘obvious’ choice for Komar’s expression yields the Newman–Unti energy. This anomalous behaviour in the radiative regime could be corrected in, at least, two ways. The first is by modifying the Komar integral according to(16) |
3.2.4 Null infinity: Angular momentum
At null infinity there is no generally accepted definition for angular momentum, and there are various, mathematically inequivalent suggestions for it. Here we review only some of those total angular momentum definitions that can be considered as the null infinity limit of some quasi-local expression, and will be discussed in the main part of the review, namely in section 9 . In their classic paper Bergmann and Thomson [58] raise the idea that while the gravitational energy-momentum is connected with the spacetime diffeomorphisms, the angular momentum should be connected with its intrinsic symmetry. Thus, the angular momentum should be analogous with the spin. Based on the tetrad formalism of general relativity and following the prescription of constructing the Noether currents in Yang–Mills theories, Bramson suggested a superpotential for the six conserved currents corresponding to the internal Lorentz-symmetry [82, ?, 83] . (For another derivation of this superpotential from Møller’s Lagrangian ( 9 ) see [347] .) If , , is a normalized spinor dyad corresponding to the orthonormal frame in ( 9 ), then the integral of the spinor form of the anti-self-dual part of this superpotential on a closed orientable 2-surface is(17) |
3.3 The necessity of quasi-locality for the observables in general relativity
3.3.1 Non-locality of the gravitational energy-momentum and angular momentum
One reaction to the non-tensorial nature of the gravitational energy-momentum density expressions was to consider the whole problem ill-defined and the gravitational energy-momentum meaningless. However, the successes discussed in the previous subsection show that the global gravitational energy-momenta and angular momenta are useful notions, and hence it could also be useful to introduce them even if the spacetime is not asymptotically flat. Furthermore, the non-tensorial nature of an object does not imply that it is meaningless. For example, the Christoffel symbols are not tensorial, but they do have geometric, and hence physical content, namely the linear connection. Indeed, the connection is a non-local geometric object, connecting the fibres of the vector bundle over different points of the base manifold. Hence any expression of the connection coefficients, in particular the gravitational energy-momentum or angular momentum, must also be non-local. In fact, although the connection coefficients at a given point can be taken zero by an appropriate coordinate/gauge transformation, they cannot be transformed to zero on an open domain unless the connection is flat. Furthermore, the superpotential of many of the classical pseudotensors (e. g. of the Einstein, Bergmann, Møller’s tetrad, Landau–Lifshitz pseudotensors), being linear in the connection coefficients, can be recovered as the pull back to the spacetime manifold of various forms of a single geometric object on the linear frame bundle, namely of the Nester–Witten 2-form, along various local sections [138, 256, 336, 337] , and the expression of the pseudotensors by their superpotentials are the pull backs of the Sparling equation [329, 126, 256] . In addition, Chang, Nester and Chen [101] found a natural quasi-local Hamiltonian interpretation of each of the pseudotensorial expressions in the metric formulation of the theory (see subsection 11.3.3 ). Therefore, the pseudotensors appear to have been ‘rehabilitated’, and the gravitational energy-momentum and angular momentum are necessarily associated with extended subsets of the spacetime. This fact is a particular consequence of a more general phenomenon [309, 207] : Since the physical spacetime is the isomorphism class of the pairs instead of a single such pair, it is meaningless to speak about the ‘value of a scalar or vector field at a point ’. What could have meaning are the quantities associated with curves (the length of a curve, or the holonomy along a closed curve), 2-surfaces (e. g. the area of a closed 2-surface) etc. Thus, if we want to associate energy-momentum and angular momentum not only to the whole (necessarily asymptotically flat) spacetime, then these quantities must be associated with extended but finite subsets of the spacetime, i. e. must be quasi-local.3.3.2 Domains for quasi-local quantities
The quasi-local quantities (usually the integral of some local expression of the field variables) are associated with a certain type of subset of spacetime. In four dimensions there are three natural candidates:(18) |
(19) |
3.3.3 Strategies to construct quasi-local quantities
There are two natural ways of finding the quasi-local energy-momentum and angular momentum. The first is to follow some systematic procedure, while the second is the ‘quasi-localization’ of the global energy-momentum and angular momentum expressions. One of the two systematic procedures could be called the Lagrangian approach: The quasi-local quantities are integrals of some superpotential derived from the Lagrangian via a Noether-type analysis. The advantage of this approach could be its manifest Lorentz-covariance. On the other hand, since the Noether current is determined only through the Noether identity, which contains only the divergence of the current itself, the Noether current and its superpotential is not uniquely determined. In addition (as in any approach), a gauge reduction (for example in the form of a background metric or reference configuration) and a choice for the ‘translations’ and ‘boost-rotations’ should be made. The other systematic procedure might be called the Hamiltonian approach: At the end of a fully quasi-local (covariant or not) Hamiltonian analysis we would have a Hamiltonian, and its value on the constraint surface in the phase space yields the expected quantities. Here the main idea is that of Regge and Teitelboim [304] that the Hamiltonian must reproduce the correct field equations as the flows of the Hamiltonian vector fields, and hence, in particular, the correct Hamiltonian must be functionally differentiable with respect to the canonical variables. This differentiability may restrict the possible ‘translations’ and ‘boost-rotations’ too. However, if we are not interested in the structure of the quasi-local phase space, then, as a short-cut, we can use the Hamilton–Jacobi method to define the quasi-local quantities. The resulting expression is a 2-surface integral. Nevertheless, just as in the Lagrangian approach, this general expression is not uniquely determined, because the action can be modified by adding an (almost freely chosen) boundary term to it. Furthermore, the ‘translations’ and ‘boost-rotations’ are still to be specified. On the other hand, at least from a pragmatic point of view, the most natural strategy to introduce the quasi-local quantities would be some ‘quasi-localization’ of those expressions that gave the global energy-momentum and angular momentum of asymptotically flat spacetimes. Therefore, respecting both strategies, it is also legitimate to consider the Winicour–Tamburino-type (linkage) integrals and the charge integrals of the curvature. Since the global energy-momentum and angular momentum of asymptotically flat spacetimes can be written as 2-surface integrals at infinity (and, as we will see in subsection 7.1.1 , both the energy-momentum and angular momentum of the source in the linear approximation and the gravitational mass in the Newtonian theory of gravity can also be written as 2-surface integrals), the 2-surface observables can be expected to have special significance. Thus, to summarize, if we want to define reasonable quasi-local energy-momentum and angular momentum as 2-surface observables, then three things must be specified:4 Tools to Construct and Analyze the Quasi-Local Quantities
Having accepted that the gravitational energy-momentum and angular momentum should be introduced at the quasi-local level, we next need to discuss the special tools and concepts that are needed in practice to construct (or even to understand the various special) quasi-local expressions. Thus, first, we review the geometry of closed spacelike 2-surfaces, with special emphasis on the so-called 2-surface data. Then, in the remaining two subsections, we discuss the special situations where there is a more or less generally accepted ‘standard’ definition for the energy-momentum (or at least for the mass) and angular momentum. In these situations any reasonable quasi-local quantity should reduce to them.4.1 The geometry of spacelike 2-surfaces
The first systematic study of the geometry of spacelike 2-surfaces from the point of view of quasi-local quantities is probably due to Tod [358, 363] . Essentially, his approach is based on the GHP formalism [148] . Although this is a very effective and flexible formalism [148, 298, 299, 200, 331] , its form is not spacetime covariant. Since in many cases the covariance of a formalism itself already gives some hint how to treat and solve the problem at hand, here we concentrate mainly on a spacetime–covariant description of the geometry of the spacelike 2-surfaces, developed gradually in [339, 341, 342, 343, 143] . The emphasis will be on the geometric structures rather than the technicalities. In the last paragraph, we comment on certain objects appearing in connection with families of spacelike 2-surfaces.4.1.1 The Lorentzian vector bundle
The restriction to the closed, orientable spacelike 2-surface of the tangent bundle of the spacetime has a unique decomposition to the -orthogonal sum of the tangent bundle of and the bundle of the normals, denoted by . Then all the geometric structures of the spacetime (metric, connection, curvature) can be decomposed in this way. If and are timelike and spacelike unit normals, respectively, being orthogonal to each other, then the projection to and is and , respectively. The induced 2-metric and the corresponding area 2-form on will be denoted by and , respectively, while the area 2-form on the normal bundle will be . The bundle together with the fibre metric and the projection will be called the Lorentzian vector bundle over . For the discussion of the global topological properties of the closed orientable 2-manifolds, see for example [5] .4.1.2 Connections
The spacetime covariant derivative operator defines two covariant derivatives on . The first, denoted by , is analogous to the induced (intrinsic) covariant derivative on (one-codimensional) hypersurfaces: for any section of . Obviously, annihilates both the fibre metric and the projection . However, since for 2-surfaces in four dimensions the normal is not uniquely determined, we have the ‘boost gauge freedom’ , . The induced connection will have a nontrivial part on the normal bundle, too. The corresponding (normal part of the) connection 1-form on can be characterized, for example, by . Therefore, the connection can be considered as a connection on coming from a connection on the -principal bundle of the -orthonormal frames adapted to . The other connection, , is analogous to the Sen connection [316] , and is defined simply by . This annihilates only the fibre metric, but not the projection. The difference of the connections and turns out to be just the extrinsic curvature tensor: . Here , and and are the standard (symmetric) extrinsic curvatures corresponding to the individual normals and , respectively. The familiar expansion tensors of the future pointing outgoing and ingoing null normals, and , respectively, are and , and the corresponding shear tensors and are defined by their trace-free part. Obviously, and (and hence the expansion and shear tensors , , and ) are boost-gauge dependent quantities (and it is straightforward to derive their transformation from the definitions), but their combination is boost-gauge invariant. In particular, it defines a natural normal vector field to by , where , , and are the relevant traces. is called the main extrinsic curvature vector of . If , then their norm is , and they are orthogonal to each other: . It is easy to show that , i. e. is the uniquely pointwise determined direction orthogonal to the 2-surface in which the expansion of the surface is vanishing. If is not null, then defines an orthonormal frame in the normal bundle (see for example [7] ). If is non-zero but (e. g. future pointing) null, then there is a uniquely determined null normal to such that , and hence is a uniquely determined null frame. Therefore, the 2-surface admits a natural gauge choice in the normal bundle unless is vanishing. Geometrically, is a connection coming from a connection on the -principal fibre bundle of the -orthonormal frames. The curvature of the connections and , respectively, are(20) |
(21) |
4.1.3 Convexity conditions
To prove certain statements on quasi-local quantities various forms of the convexity of must be assumed. The convexity of in a 3-geometry is defined by the positive definiteness of its extrinsic curvature tensor. If the embedding space is flat, then by the Gauss equation this is equivalent to the positivity of the scalar curvature of the intrinsic metric of . If is in a Lorentzian spacetime then the weakest convexity conditions are conditions only on the mean null curvatures: will be called weakly future convex if the outgoing null normals are expanding on , i. e. , and weakly past convex if [363] . is called mean convex [177] if on , or, equivalently, if is timelike. To formulate stronger convexity conditions we must consider the determinant of the null expansions and . Note that although the expansion tensors, and in particular the functions , , and are gauge dependent, their sign is gauge invariant. Then will be called future convex if and , and past convex if and [363, 342] . A different kind of convexity condition, based on global concepts, will be used in subsection 6.1.3 .4.1.4 The spinor bundle
The connections and determine connections on the pull back to of the bundle of unprimed spinors. The natural decomposition defines a chirality on the spinor bundle in the form of the spinor , which is analogous to the matrix in the theory of Dirac spinors. Then the extrinsic curvature tensor above is a simple expression of and (and their complex conjugate), and the two covariant derivatives on are related to each other by . The curvature of can be expressed by the curvature of , the spinor and its -derivative. We can form the scalar invariants of the curvatures according to(22) |
(23) |
4.1.5 Curvature identities
The complete decomposition of into its irreducible parts gives , the Dirac–Witten operator, and , the 2-surface twistor operator. A Sen–Witten type identity for these irreducible parts can be derived. Taking its integral one has(24) |
4.1.6 The GHP formalism
A GHP spin frame on the 2-surface is a normalized spinor basis , , such that the complex null vectors and are tangent to (or, equivalently, the future pointing null vectors and are orthogonal to ). Note, however, that in general a GHP spin frame can be specified only locally, but not globally on the whole . This fact is connected with the non-triviality of the tangent bundle of the 2-surface. For example, on the 2-sphere every continuous tangent vector field must have a zero, and hence, in particular, the vectors and cannot form a globally defined basis on . Consequently, the GHP spin frame cannot be globally defined either. The only closed orientable 2-surface with globally trivial tangent bundle is the torus. Fixing a GHP spin frame on some , the components of the spinor and tensor fields on will be local representatives of cross sections of appropriate complex line bundles of scalars of type [148, 298] : A scalar is said to be of type if under the rescaling , of the GHP spin frame with some nowhere vanishing complex function the scalar transforms as . For example , , and are of type , , and , respectively. The components of the Weyl and Ricci spinors, , , , . . . , , , . . . , etc., also have definite type. In particular, has type . A global section of is a collection of local cross sections such that forms a covering of and on the non-empty overlappings, e. g. on the local sections are related to each other by , where is the transition function between the GHP spin frames: and . The connection defines a connection on the line bundles [148, 298] . The usual edth operators, and , are just the directional derivatives and on the domain of the GHP spin frame . These locally defined operators yield globally defined differential operators, denoted also by and , on the global sections of . It might be worth emphasizing that the GHP spin coefficients and , which do not have definite -type, play the role of the two components of the connection 1-form, and they are built both from the connection 1-form for the intrinsic Riemannian geometry of and the connection 1-form in the normal bundle. and are elliptic differential operators, thus their global properties, e. g. the dimension of their kernel, are connected with the global topology of the line bundle they act on, and, in particular, with the global topology of . These properties are discussed in [143] for general, and in [128, 42, 340] for spherical topology.4.1.7 Irreducible parts of the derivative operators
Using the projection operators , the irreducible parts and can be decomposed further into their right handed and left handed parts. In the GHP formalism these chiral irreducible parts are(25) |
4.1.8 -connection 1-form versus anholonomicity
Obviously, all the structures we have considered can be introduced on the individual surfaces of oneor two-parameter families of surfaces, too. In particular [176] , let the 2-surface be considered as the intersection of the null hypersurfaces formed, respectively, by the outgoing and the ingoing light rays orthogonal to , and let the spacetime (or at least a neighbourhood of ) be foliated by two one-parameter families of smooth hypersurfaces and , where , such that and . One can form the two normals, , which are null on and , respectively. Then we can define , for which , where . (If is chosen to be 1 on , then is precisely the connection 1-form above.) Then the so-called anholonomicity is defined by . Since is invariant with respect to the rescalings and of the functions defining the foliations by those functions which preserve , it was claimed in [176] that depends only on . However, this implies only that is invariant with respect to a restricted class of the change of the foliations, and that is invariantly defined only by this class of the foliations rather than the 2-surface. In fact, does depend on the foliation: Starting with a different foliation defined by the functions and for some , the corresponding anholonomicity would also be invariant with respect to the restricted changes of the foliations above, but the two anholonomicities, and , would be different: . Therefore, the anholonomicity is still a gauge dependent quantity.4.2 Standard situations to evaluate the quasi-local quantities
There are exact solutions to the Einstein equations and classes of special (e. g. asymptotically flat) spacetimes in which there is a commonly accepted definition of energy-momentum (or at least mass) and angular momentum. In this subsection we review these situations and recall the definition of these ‘standard’ expressions.4.2.1 Round spheres
If the spacetime is spherically symmetric, then a 2-sphere which is a transitivity surface of the rotation group is called a round sphere. Then in a spherical coordinate system the spacetime metric takes the form , where and are functions of and . (Hence is the so-called area-coordinate). Then with the notations of subsection 4.1 , one obtains . Based on the investigations of Misner, Sharp and Hernandez [258, 193] , Cahill and McVitte [95] found(26) |
(27) |
can be thought of as the 0-component of some quasi-local energy-momentum 4-vector, but, just because of the spherical symmetry, its spatial parts are vanishing. Thus can also be interpreted as the mass, the length of this energy-momentum 4-vector.
4.2.2 Small surfaces
In the literature there are two notions of small surfaces, the first is that of the small spheres (both in the light cone of a point and in a spacelike hypersurface), introduced first by Horowitz and Schmidt [198] , and the other is the concept of the small ellipsoids in some spacelike hypersurface, considered first by Woodhouse in [229] . A small sphere in the light cone is a cut of the future null cone in the spacetime by a spacelike hypersurface, and the geometry of the sphere is characterized by data at the vertex of the cone. The sphere in a hypersurface consists of those points of a given spacelike hypersurface, whose geodesic distance in the hypersurface from a given point , the centre, is a small given value, and the geometry of this sphere is characterized by data at this centre. Small ellipsoids are 2-surfaces in a spacelike hypersurface with a more general shape. To define the first, let be a point, and a future directed unit timelike vector at . Let , the ‘future null cone of in ’ (i. e. the boundary of the chronological future of ). Let be the future pointing null tangent to the null geodesic generators of such that, at the vertex , . With this condition we fix the scale of the affine parameter on the different generators, and hence by requiring we fix the parameterization completely. Then, in an open neighbourhood of the vertex , is a smooth null hypersurface, and hence for sufficiently small the set is a smooth spacelike 2-surface and homeomorphic to . is called a small sphere of radius with vertex . Note that the condition fixes the boost gauge. Completing to a Newman–Penrose complex null tetrad such that the complex null vectors and are tangent to the 2-surfaces , the components of the metric and the spin coefficients with respect to this basis can be expanded as series in . Then the GHP equations can be solved with any prescribed accuracy for the expansion coefficients of the metric on , the GHP spin coefficients , , , , and and the higher order expansion coefficients of the curvature in terms of the lower order curvature components at . Hence the expression of any quasi-local quantity for the small sphere can be expressed as a series of , where the expansion coefficients are still functions of the coordinates, or , on the unit sphere . If the quasi-local quantity is spacetime–covariant, then the unit sphere integrals of the expansion coefficients must be spacetime covariant expressions of the metric and its derivatives up to some finite order at and the ‘time axis’ . The necessary degree of the accuracy of the solution of the GHP equations depends on the nature of and on whether the spacetime is Ricci-flat in a neighbourhood of or not. These solutions of the GHP equations, with increasing accuracy, are given in [198, 229, 91, 344] . Obviously, we can calculate the small sphere limit of various quasi-local quantities built from the matter fields in the Minkowski spacetime too. In particular [344] , the small sphere expressions for the quasi-local energy-momentum and the (anti-self-dual part of the) quasi-local angular momentum of the matter fields based on , respectively, are(28) |
(29) |
If, in addition, the spinor constituent of is required to be parallel propagated along , then the tetrad becomes completely fixed, yielding the vanishing of several (combinations of the) spin coefficients.
As we will see soon, the leading term of the small sphere expression of the energy-momenta in non-vacuum is of order , in vacuum it is , while that of the angular momentum is and , respectively.
4.2.3 Large spheres near the spatial infinity
Near spatial infinity we have the a priori and fall-off for the 3-metric and extrinsic curvature , respectively, and both the evolution equations of general relativity and the conservation equation for the matter fields preserve these conditions. The spheres of coordinate radius in are called large spheres if the values of are large enough such that the asymptotic expansions of the metric and extrinsic curvature are legitimate. Introducing some coordinate system, e. g. the complex stereographic coordinates, on one sphere and then extending that to the whole along the normals of the spheres, we obtain a coordinate system on . Let , , be a GHP spinor dyad on adapted to the large spheres in such a way that and are tangent to the spheres and , the future directed unit normal of . These conditions fix the spinor dyad completely, and, in particular, , the outward directed unit normal to the spheres tangent to . The fall-off conditions yield that the spin coefficients tend to their flat spacetime value like and the curvature components to zero like . Expanding the spin coefficients and curvature components as power series of , one can solve the field equations asymptotically (see [47, 43] for a different formalism). However, in most calculations of the large sphere limit of the quasi-local quantities only the leading terms of the spin coefficients and curvature components appear. Thus it is not necessary to solve the field equations for their second or higher order non-trivial expansion coefficients. Using the flat background metric and the corresponding derivative operator we can define a spinor field to be constant if . Obviously, the constant spinors form a two complex dimensional vector space. Then by the fall-off properties . Hence we can define the asymptotically constant spinor fields to be those that satisfy , where is the intrinsic Levi-Civita derivative operator. Note that this implies that, with the notations of ( 25 ), all the chiral irreducible parts, , , and , of the derivative of the asymptotically constant spinor field are .Because of the fall-off, no essential ambiguity in the definition of the large spheres arises from the use of the coordinate radius instead of the physical radial distance.
4.2.4 Large spheres near null infinity
Let the spacetime be asymptotically flat at future null infinity in the sense of Penrose [286, 287, 288, 299] (see also [147] ), i. e. the physical spacetime can be conformally compactified by an appropriate boundary . Then future null infinity will be a null hypersurface in the conformally rescaled spacetime. Topologically it is , and the conformal factor can always be chosen such that the induced metric on the compact spacelike slices of is the metric of the unit sphere. Fixing such a slice (called ‘the origin cut of ’) the points of can be labelled by a null coordinate, namely the affine parameter along the null geodesic generators of measured from and, for example, the familiar complex stereographic coordinates , defined first on the unit sphere and then extended in a natural way along the null generators to the whole . Then any other cut of can be specified by a function . In particular, the cuts are obtained from by a pure time translation. The coordinates can be extended to an open neighbourhood of in the spacetime in the following way. Let be the family of smooth outgoing null hypersurfaces in a neighbourhood of such that they intersect the null infinity just in the cuts , i. e. . Then let be the affine parameter in the physical metric along the null geodesic generators of . Then forms a coordinate system. The , 2-surfaces (or simply if no confusion can arise) are spacelike topological 2-spheres, which are called large spheres of radius near future null infinity. Obviously, the affine parameter is not unique, its origin can be changed freely: is an equally good affine parameter for any smooth . Imposing certain additional conditions to rule out such coordinate ambiguities we arrive at a ‘Bondi type coordinate system’. In many of the large sphere calculations of the quasi-local quantities the large spheres should be assumed to be large spheres not only in a general null, but in a Bondi type coordinate system. For the detailed discussion of the coordinate freedom left at the various stages in the introduction of these coordinate systems, see for example [277, 276, 82] . In addition to the coordinate system we need a Newman–Penrose null tetrad, or rather a GHP spinor dyad, , , on the hypersurfaces . (Thus boldface indices are referring to the GHP spin frame.) It is natural to choose such that be the tangent of the null geodesic generators of , and itself be constant along . Newman and Unti [277] chose to be parallel propagated along . This choice yields the vanishing of a number of spin coefficients (see for example the review [276] ). The asymptotic solution of the Einstein–Maxwell equations as a series of in this coordinate and tetrad system is given in [277, 130, 298] , where all the non-vanishing spin coefficients and metric and curvature components are listed. In this formalism the gravitational waves are represented by the -derivative of the asymptotic shear of the null geodesic generators of the outgoing null hypersurfaces . From the point of view of the large sphere calculations of the quasi-local quantities the choice of Newman and Unti for the spinor basis is not very convenient. It is more natural to adapt the GHP spin frame to the family of the large spheres of constant ‘radius’ , i. e. to require and to be tangents of the spheres. This can be achieved by an appropriate null rotation of the Newman–Unti basis about the spinor . This rotation yields a change of the spin coefficients and the metric and curvature components. As far as the present author is aware of, this rotation with the highest accuracy was done for the solutions of the Einstein–Maxwell system by Shaw [323] . In contrast to the spatial infinity case, the ‘natural’ definition of the asymptotically constant spinor fields yields identically zero spinors in general [81] . Nontrivial constant spinors in this sense could exist only in the absence of the outgoing gravitational radiation, i. e. when . In the language of subsection 4.1.7 , this definition would be , , and . However, as Bramson showed [81] , half of these conditions can be imposed. Namely, at future null infinity (and at past null infinity ) can always be imposed asymptotically, and it has two linearly independent solutions , , on (or on , respectively). The space of its solutions turns out to have a natural symplectic metric , and we refer to as future asymptotic spin space. Its elements are called asymptotic spinors, and the equations the future/past asymptotic twistor equations. At asymptotic spinors are the spinor constituents of the BMS translations: Any such translation is of the form for some constant Hermitian matrix . Similarly, (apart from the proper supertranslation content) the components of the anti-self-dual part of the boost-rotation BMS vector fields are , where are the standard Pauli matrices (divided by ) [347] . Asymptotic spinors can be recovered as the elements of the kernel of several other operators built from , , and too. In the present review we use only the fact that asymptotic spinors can be introduced as anti-holomorphic spinors (see also subsection 8.2.1 ), i. e. the solutions of (and at past null infinity as holomorphic spinors), and as special solutions of the 2-surface twistor equation (see also subsection 7.2.1 ). These operators, together with others reproducing the asymptotic spinors, are discussed in [347] . The Bondi–Sachs energy-momentum given in the Newman–Penrose formalism has already become its ‘standard’ form. It is the unit sphere integral on the cut of a combination of the leading term of the Weyl spinor component , the asymptotic shear and its -derivative, weighted by the first four spherical harmonics (see for example [276, 299] ):(30) |
(31) |
In the so-called Bondi coordinate system the radial coordinate is the luminosity distance , which tends to the affine parameter asymptotically.
4.2.5 Other special situations
In the weak field approximation of general relativity [365, 21, 369, 299, 221] the gravitational field is described by a symmetric tensor field on Minkowski spacetime , and the dynamics of the field is governed by the linearized Einstein equations, i. e. essentially the wave equation. Therefore, the tools and techniques of the Poincare-invariant field theories, in particular the Noether–Belinfante–Rosenfeld procedure outlined in subsection 2.1 and the ten Killing vectors of the background Minkowski spacetime, can be used to construct the conserved quantities. It turns out that the symmetric energy-momentum tensor of the field is essentially the second order term in the Einstein tensor of the metric . Thus in the linear approximation the field does not contribute to the global energy-momentum and angular momentum of the matter+gravity system, and hence these quantities have the form ( 5 ) with the linearized energy-momentum tensor of the matter fields. However, as we will see in subsection 7.1.1 , this energy-momentum and angular momentum can be re-expressed as a charge integral of the (linearized) curvature [333, 200, 299] . pp-waves spacetimes are defined to be those that admit a constant null vector field , and they are interpreted as describing pure plane-fronted gravitational waves with parallel rays. If matter is present then it is necessarily pure radiation with wavevector , i. e. holds [236] . A remarkable feature of the pp-wave metrics is that, in the usual coordinate system, the Einstein equations become a two dimensional linear equation for a single function. In contrast to the approach adopted almost exclusively, Aichelburg [3] considered this field equation as an equation for a boundary value problem. As we will see, from the point of view of the quasi-local observables this is a particularly useful and natural standpoint. If a pp-wave spacetime admits an additional spacelike Killing vector with closed orbits, i. e. it is cyclically symmetric too, then and are necessarily commuting and are orthogonal to each other, because otherwise an additional timelike Killing vector would also be admitted [335] . Since the final state of stellar evolution (the neutron star or the black hole state) is expected to be described by an asymptotically flat stationary, axis-symmetric spacetime, the significance of these spacetimes is obvious. It is conjectured that this final state is described by the Kerr–Newman (either outer or black hole) solution with some well defined mass, angular momentum and electric charge parameters [369] . Thus axis-symmetric 2-surfaces in these solutions may provide domains which are general enough but for which the quasi-local quantities are still computable. According to a conjecture by Penrose [291] , the (square root of the) area of the event horizon provides a lower bound for the total ADM energy. For the Kerr–Newman black hole this area is . Thus, particularly interesting 2-surfaces in these spacetimes are the spacelike cross sections of the event horizon [60] . There is a well defined notion of total energy-momentum not only in the asymptotically flat, but even in the asymptotically anti-de-Sitter spacetimes too. This is the Abbott–Deser energy [1] , whose positivity has also been proven under similar conditions that we had to impose in the positivity proof of the ADM energy [157] . The conformal technique, initiated by Penrose, is used to give a precise definition of the asymptotically anti-de-Sitter spacetimes and to study their general, basic properties in [26] . A comparison and analysis of the various definitions of mass for asymptotically anti-de-Sitter metrics is given in [114] . Thus it is natural to ask whether a specific quasi-local energy-momentum expression is able to reproduce the Abbott–Deser energy-momentum in this limit or not.4.3 On lists of criteria of reasonableness of the quasi-local quantities
In the literature there are various, more or less ad hoc, ‘lists of criteria of reasonableness’ of the quasi-local quantities (see for example [127, 108] ). However, before discussing them, it seems useful to formulate first some general principles that any quasi-local quantity should satisfy.4.3.1 General expectations
In non-gravitational physics the notions of conserved quantities are connected with symmetries of the system, and they are introduced through some systematic procedure in the Lagrangian and/or Hamiltonian formalism. In general relativity the total energy-momentum and angular momentum are 2-surface observables, thus we concentrate on them even at the quasi-local level. These facts motivate our three a priori expectations:(32) |
4.3.2 Pragmatic criteria
Since in certain special situations there are generally accepted definitions for the energy-momentum and angular momentum, it seems reasonable to expect that in these situations the quasi-local quantities reduce to them. One half of the pragmatic criteria is just this expectation, and the other is a list of some a priori requirements on the behaviour of the quasi-local quantities. One such list for the energy-momentum and mass, based mostly on [127, 108] and the properties of the quasi-local energy-momentum of the matter fields of subsection 2.2 , might be the following:4.3.3 Incompatibility of certain ‘natural’ expectations
As Eardley noted in [127] , probably no quasi-local energy definition exists which would satisfy all of his criteria. In fact, it is easy to see that this is the case. Namely, any quasi-local energy definition which reduces to the ‘standard’ expression for round spheres cannot be monotonic, as the closed Friedmann–Robertson–Walker or the spacetimes show explicitly. The points where the monotonicity breaks down are the extremal (maximal or minimal) surfaces, which represent event horizon in the spacetime. Thus one may argue that since the event horizon hides a portion of spacetime, we cannot know the details of the physical state of the matter+gravity system behind the horizon. Hence, in particular, the monotonicity of the quasi-local mass may be expected to break down at the event horizon. However, although for stationary systems (or at the moment of time symmetry of a time symmetric system) the event horizon corresponds to an apparent horizon (or to an extremal surface, respectively), for general non-stationary systems the concepts of the event and apparent horizons deviate. Thus the causal argument above does not seem possible to be formulated in the hypersurface of subsection 4.3.2 . Actually, the root of the non-monotonicity is the fact that the quasi-local energy is a 2-surface observable in the sense of 1 in subsection 4.3.1 above. This does not mean, of course, that in certain restricted situations the monotonicity (‘local monotonicity’) could not be proven. This local monotonicity may be based, for example, on Lie dragging of the 2-surface along some special spacetime vector field. On the other hand, in the literature sometimes the positivity and the monotonicity requirements are confused, and there is an ‘argument’ that the quasi-local gravitational energy cannot be positive definite, because the total energy of the closed universes must be zero. However, this argument is based on the implicit assumption that the quasi-local energy is associated with a compact three dimensional domain, which, together with the positive definiteness requirement would, in fact, imply the monotonicity and a positive total energy for the closed universe. If, on the other hand, the quasi-local energy-momentum is associated with 2-surfaces, then the energy may be positive definite and not monotonic. The standard round sphere energy expression ( 26 ) in the closed Friedmann–Robertson–Walker spacetime, or, more generally, the Dougan–Mason energy-momentum (see subsection 8.2.3 ) are such examples.5 The Bartnik Mass and its Modifications
5.1 The Bartnik mass
5.1.1 The main idea
One of the most natural ideas of quasi-localization of the familiar ADM mass is due to Bartnik [38, 37] . His idea is based on the positivity of the ADM energy, and, roughly, can be summarized as follows. Let be a compact, connected 3-manifold with connected boundary , and let be a (negative definite) metric and a symmetric tensor field on such that they, as an initial data set, satisfy the dominant energy condition: If and , then . For the sake of simplicity we denote the triple by . Then let us consider all the possible asymptotically flat initial data sets with a single asymptotic end, denoted simply by , which satisfy the dominant energy condition, have finite ADM energy and are extensions of above through its boundary . The set of these extensions will be denoted by . By the positive energy theorem has non-negative ADM energy , which is zero precisely when is a data set for the flat spacetime. Then we can consider the infimum of the ADM energies, , where the infimum is taken on . Obviously, by the non-negativity of the ADM energies this infimum exists and is non-negative, and it is tempting to define the quasi-local mass of by this infimum. However, it is easy to see that, without further conditions on the extensions of , this infimum is zero. In fact, can be extended to an asymptotically flat initial data set with arbitrarily small ADM energy such that contains a horizon (for example in the form of an apparent horizon) between the asymptotically flat end and . In particular, in the ‘ -spacetime’, discussed in subsection 4.2.1 on the round spheres, the spherically symmetric domain bounded by the maximal surface (with arbitrarily large round-sphere mass ) has an asymptotically flat extension, the -spacetime itself, with arbitrarily small ADM mass . Obviously, the fact that the ADM energies of the extensions can be arbitrarily small is a consequence of the presence of a horizon hiding from the outside. This led Bartnik [38, 37] to formulate his suggestion for the quasi-local mass of . He concentrated on the time-symmetric data sets (i. e. those for which the extrinsic curvature is vanishing), when the horizon appears to be a minimal surface of topology in (see for example [152] ), and the dominant energy condition is just the requirement of the non-negativity of the scalar curvature: . Thus, if denotes the set of asymptotically flat Riemannian geometries with non-negative scalar curvature and finite ADM energy that contain no stable minimal surface, then Bartnik’s mass is(33) |
(34) |
Since we take the infimum, we could equally take the ADM masses, which are the minimum values of the zero-th component of the energy-momentum four-vectors in the different Lorentz frames, instead of the energies.
5.1.2 The main properties of
The first immediate consequence of ( 33 ) is the monotonicity of the Bartnik mass: If , then , and hence . Obviously, by definition ( 33 ) one has for any . Thus if is any quasi-local mass functional which is larger than (i. e. which assigns a non-negative real to any such that for any allowed ), furthermore if for any , then by the definition of the infimum in ( 33 ) one has for any . Therefore, is the largest mass functional satisfying for any . Another interesting consequence of the definition of , due to W. Simon, is that if is any asymptotically flat, time symmetric extension of with non-negative scalar curvature satisfying , then there is a black hole in in the form of a minimal surface between and the infinity of (see for example [40] ). As we saw, the Bartnik mass is non-negative, and, obviously, if is flat (and hence is a data set for the flat spacetime), then . The converse of this statement is also true [202] : If , then is locally flat. The Bartnik mass tends to the ADM mass [202] : If is an asymptotically flat Riemannian 3-geometry with non-negative scalar curvature and finite ADM mass , and if , , is a sequence of solid balls of coordinate radius in , then . The proof of these two results is based on the use of the Hawking energy (see subsection 6.1 ), by means of which a positive lower bound for can be given near the non-flat points of . In the proof of the second statement one must use the fact that the Hawking energy tends to the ADM energy, which, in the time symmetric case, is just the ADM mass. The proof that the Bartnik mass reduces to the ‘standard expression’ for round spheres is a nice application of the Riemannian Penrose inequality [202] : Let be a spherically symmetric Riemannian 3-geometry with spherically symmetric boundary . One can form its ‘standard’ round-sphere energy (see subsection 4.2.1 ), and take its spherically symmetric asymptotically flat vacuum extension (see [38, 40] ). By the Birkhoff theorem the exterior part of is a part of a hypersurface of the vacuum Schwarzschild solution, and its ADM mass is just . Then any asymptotically flat extension of can also be considered as (a part of) an asymptotically flat time symmetric hypersurface with minimal surface, whose area is . Thus by the Riemannian Penrose inequality [202] . Therefore, the Bartnik mass of is just the ‘standard’ round sphere expression .5.1.3 The computability of the Bartnik mass
Since for any given the set of its extensions is a huge set, it is almost hopeless to parameterize it. Thus, by the very definition, it seems very difficult to compute the Bartnik mass for a given, specific . Without some computational method the potentially useful properties of would be lost from the working relativist’s arsenal. Such a computational method might be based on a conjecture of Bartnik [38, 40] : The infimum in definition ( 33 ) of the mass is realized by an extension of such that the exterior region, , is static, the metric is Lipschitz-continuous across the 2-surface , and the mean curvatures of of the two sides are equal. Therefore, to compute for a given , one should find an asymptotically flat, static vacuum metric satisfying the matching conditions on , and the Bartnik mass is the ADM mass of . As Corvino showed [115] , if there is an allowed extension of for which , then the extension is static; furthermore, if , and has an allowed extension for which , then is static. Thus the proof of Bartnik’s conjecture is equivalent to the proof of the existence of such an allowed extension. The existence of such an extension is proven in [257] for geometries close enough to the Euclidean one and satisfying a certain reflection symmetry, but the general existence proof is still lacking. Bartnik’s conjecture is that determines this exterior metric uniquely [40] . He conjectures [38, 40] that a similar computation method can be found for the mass , defined in ( 34 ), too, where the exterior metric should be stationary. This second conjecture is also supported by partial results [116] : If is any compact vacuum data set, then it has an asymptotically flat vacuum extension which is a spacelike slice of a Kerr spacetime outside a large sphere near spatial infinity. To estimate one can construct admissible extensions of in the form of the metrics in quasi-spherical form [39] . If the boundary is a metric sphere of radius with non-negative mean curvature , then can be estimated from above in terms of and .5.2 Bray’s modifications
Another, slightly modified definition for the quasi-local mass was suggested by Bray [84, 87] . Here we summarize his ideas. Let be any asymptotically flat initial data set with finitely many asymptotic ends and finite ADM masses, and suppose that the dominant energy condition is satisfied on . Let be any fixed 2-surface in which encloses all the asymptotic ends except one, say the -th (i. e. let be homologous to a large sphere in the -th asymptotic end). The outside region with respect to , denoted by , will be the subset of containing the -th asymptotic end and bounded by , while the inside region, , is the (closure of) . Next Bray defines the ‘extension’ of by replacing by a smooth asymptotically flat end of any data set satisfying the dominant energy condition. Similarly, the ‘fill-in’ of is obtained from by replacing by a smooth asymptotically flat end of any data set satisfying the dominant energy condition. Finally, the surface will be called outer-minimizing if for any closed 2-surface enclosing one has . Let be outer-minimizing, and let denote the set of extensions of in which is still outer-minimizing, and denote the set of fill-ins of . If and denotes the infimum of the area of the 2-surfaces enclosing all the ends of except the outer one, then Bray defines the outer and inner mass, and , respectively, by6 The Hawking Energy and its Modifications
6.1 The Hawking energy
6.1.1 The definition
Studying the perturbation of the dust-filled Friedmann–Robertson–Walker spacetimes, Hawking found that(35) |
6.1.2 The Hawking energy for spheres
Obviously, for round spheres reduces to the standard expression ( 26 ). This implies, in particular, that the Hawking energy is not monotonic in general. Since for a Killing horizon (e. g. for a stationary event horizon) , the Hawking energy of its spacelike spherical cross sections is . In particular, for the event horizon of a Kerr–Newman black hole it is just the familiar irreducible mass . For a small sphere of radius with centre in non-vacuum spacetimes it is , while in vacuum it is , where is the energy-momentum tensor and is the Bel–Robinson tensor at [198] . The first result shows that in the lowest order the gravitational ‘field’ does not have a contribution to the Hawking energy, that is due exclusively to the matter fields. Thus in vacuum the leading order of must be higher than . Then even a simple dimensional analysis shows that the number of the derivatives of the metric in the coefficient of the order term in the power series expansion of is . However, there are no tensorial quantities built from the metric and its derivatives such that the total number of the derivatives involved would be three. Therefore, in vacuum, the leading term is necessarily of order , and its coefficient must be a quadratic expression of the curvature tensor. It is remarkable that for small spheres is positive definite both in non-vacuum (provided the matter fields satisfy, for example, the dominant energy condition) and vacuum. This shows, in particular, that should be interpreted as energy rather than as mass: For small spheres in a pp-wave spacetime is positive, while, as we saw this for the matter fields in subsection 2.2.3 , a mass expression could be expected to be zero. (We will see in subsections 8.2.3 and 13.5 that, for the Dougan–Mason energy-momentum, the vanishing of the mass characterizes the pp-wave metrics completely.) Using the second expression in ( 35 ) it is easy to see that at future null infinity tends to the Bondi–Sachs energy. A detailed discussion of the asymptotic properties of near null infinity, both for radiative and stationary spacetimes is given in [323, 325] . Similarly, calculating for large spheres near spatial infinity in an asymptotically flat spacelike hypersurface, one can show that it tends to the ADM energy.6.1.3 Positivity and monotonicity properties
In general the Hawking energy may be negative, even in the Minkowski spacetime. Geometrically this should be clear, since for an appropriately general (e. g. concave) 2-surface the integral could be less than . Indeed, in flat spacetime is proportional to by the Gauss equation. For topologically spherical 2-surfaces in the spacelike hyperplane of Minkowski spacetime is real and non-positive, and it is zero precisely for metric spheres; while for 2-surfaces in the timelike cylinder is real and non-negative, and it is zero precisely for metric spheres. If, however, is ‘round enough’ (not to be confused with the round spheres in subsection 4.2.1 , which is some form of a convexity condition, then behaves nicely [108] : will be called round enough if it is a submanifold of a spacelike hypersurface , and if among the two dimensional surfaces in which enclose the same volume as does, has the smallest area. Then it is proven by Christodoulou and Yau [108] that if is round enough in a maximal spacelike slice on which the energy density of the matter fields is non-negative (for example if the dominant energy condition is satisfied), then the Hawking energy is non-negative. Although the Hawking energy is not monotonic in general, it has interesting monotonicity properties for special families of 2-surfaces. Hawking considered one-parameter families of spacelike 2-surfaces foliating the outgoing and the ingoing null hypersurfaces, and calculated the change of [166] . These calculations were refined by Eardley [127] . Starting with a weakly future convex 2-surface and using the boost gauge freedom, he introduced a special family of spacelike 2-surfaces in the outgoing null hypersurface , where will be the luminosity distance along the outgoing null generators. He showed that is non-decreasing with , provided the dominant energy condition holds on . Similarly, for weakly past convex and the analogous family of surfaces in the ingoing null hypersurface is non-increasing. Eardley also considered a special spacelike hypersurface, filled by a family of 2-surfaces, for which is non-decreasing. By relaxing the normalization condition for the two null normals to for some , Hayward obtained a flexible enough formalism to introduce a double-null foliation (see subsection 11.2 below) of a whole neighbourhood of a mean convex 2-surface by special mean convex 2-surfaces [177] . (For the more general GHP formalism in which is not fixed, see [298] .) Assuming that the dominant energy condition holds, he showed that the Hawking energy of these 2-surfaces is non-decreasing in the outgoing, and non-increasing in the ingoing direction. In contrast to the special foliations of the null hypersurfaces above, Frauendiener defined a special spacelike vector field, the inverse mean curvature vector in the spacetime [141] . If is a weakly future and past convex 2-surface, then is an outward directed spacelike normal to . Here is the trace of the extrinsic curvature tensor: (see subsection 4.1.2 ). Starting with a single weakly future and past convex 2-surface, Frauendiener gives an argument for the construction of a one-parameter family of 2-surfaces being Lie-dragged along its own inverse mean curvature vector . Hence this family of surfaces would be analogous to the solution of the geodesic equation, where the initial point and direction in that point specify the whole solution, at least locally. Assuming that such a family of surfaces (and hence the vector field on the 3-submanifold swept by ) exists, Frauendiener showed that the Hawking energy is non-decreasing along the vector field if the dominant energy condition is satisfied. However, no investigation has been made to prove the existence of such a family of surfaces. Motivated by this result, Malec, Mars and Simon [251] considered spacelike hypersurfaces with an inverse mean curvature flow of Geroch thereon (see subsection 6.2.2 ). They showed that if the dominant energy condition and certain additional (essentially technical) assumptions hold, then the Hawking energy is monotonic. These two results are the natural adaptations for the Hawking energy of the corresponding results known for some time for the Geroch energy, aiming to prove the Penrose inequality. We return to this latter issue in subsection 13.2 only for a very brief summary.I thank Paul Tod for pointing out this to me.
6.1.4 Two generalizations
Hawking defined not only energy, but spatial momentum as well, completely analogously to how the spatial components of the Bondi–Sachs energy-momentum are related to the Bondi energy:(36) |
(37) |
6.2 The Geroch energy
6.2.1 The definition
Suppose that the 2-surface for which is defined is embedded in the spacelike hypersurface . Let be the extrinsic curvature of in and the extrinsic curvature of in . (In subsection 4.1.2 we denoted the latter by .) Then , by means of which(38) |
6.2.2 Monotonicity properties
The Geroch energy has interesting positivity and monotonicity properties along a special flow in [146, 213] . This flow is the so-called inverse mean curvature flow defined as follows. Let be a smooth function such that6.3 The Hayward energy
We saw that can be non-zero even in the Minkowski spacetime. This may motivate considering the following expression7 Penrose’s Quasi-Local Energy-Momentum and Angular Momentum
The construction of Penrose is based on twistor-theoretical ideas, and motivated by the linearized gravity integrals for energy-momentum and angular momentum. Since, however, twistor-theoretical ideas and basic notions are still considered to be some ‘special knowledge’, the review of the basic idea behind the Penrose construction is slightly more detailed than that of the others. The basic references of the field are the volumes [298, 299] by Penrose and Rindler on ‘Spinors and Spacetime’, especially volume 2, the very well readable book by Hugget and Tod [200] and the comprehensive review article [360] by Tod.7.1 Motivations
7.1.1 How do the twistors emerge?
In the Newtonian theory of gravity the mass contained in some finite 3-volume can be expressed as the flux integral of the gravitational field strength on the boundary :(39) |
(40) |
(41) |
(42) |
(43) |
(44) |
The analogous calculations using tensor methods and the real instead of spinors and the anti-self-dual (or, shortly, a.s.d.) part of would be technically more complicated [293, 294, 299, 160] .
7.1.2 Twistor space and the kinematical twistor
Recall that the space of the contravariant valence 1 twistors of Minkowski spacetime is the set of the pairs of spinor fields, which solve the so-called valence 1 twistor equation . If is a solution of this equation, then is a solution of the corresponding equation in the conformally rescaled spacetime, where and is the conformal factor. In general the twistor equation has only the trivial solution, but in the (conformal) Minkowski spacetime it has a four complex parameter family of solutions. Its general solution in the Minkowski spacetime is , where and are constant spinors. Thus the space of valence 1 twistors, the so-called twistor-space, is four complex dimensional, and hence has the structure . admits a natural Hermitian scalar product: If is another twistor, then . Its signature is , it is conformally invariant: , and it is constant on Minkowski spacetime. The metric defines a natural isomorphism between the complex conjugate twistor space, , and the dual twistor space, , by . This makes it possible to use only twistors with unprimed indices. In particular, the complex conjugate of the covariant valence 2 twistor can be represented by the so-called conjugate twistor . We should mention two special, higher valence twistors. The first is the so-called infinity twistor. This and its conjugate are given explicitly by(45) |
(46) |
(47) |
(48) |
(49) |
(50) |
7.2 The original construction for curved spacetimes
7.2.1 2-surface twistors and the kinematical twistor
In general spacetimes the twistor equations have only the trivial solution. Thus to be able to associate a kinematical twistor to a closed orientable spacelike 2-surface in general, the conditions on the spinor field had to be relaxed. Penrose’s suggestion [293, 294] is to consider in ( 40 ) to be the symmetrized product of spinor fields that are solutions of the ‘tangential projection to ’ of the valence 1 twistor equation, the so-called 2-surface twistor equation. (The equation obtained as the ‘tangential projection to ’ of the valence 2 twistor equation ( 43 ) would be under-determined [294] .) Thus the quasi-local quantities are searched for in the form of a charge integral of the curvature:(51) |
7.2.2 The Hamiltonian interpretation of the kinematical twistor
For the solutions and of the 2-surface twistor equation the spinor identity ( 24 ) reduces to Tod’s expression [293, 299, 360] for the kinematical twistor, making it possible to re-express by the integral of the Nester–Witten 2-form [340] . Indeed, if(52) |
7.2.3 The Hermitian scalar product and the infinity twistor
In general the natural pointwise Hermitian scalar product, defined by , is not constant on , thus it does not define a Hermitian scalar product on the 2-surface twistor space. As is shown in [214, 217, 358] , is constant on for any two 2-surface twistors if and only if can be embedded, at least locally, into some conformal Minkowski spacetime with its intrinsic metric and extrinsic curvatures. Such 2-surfaces are called non-contorted, while those that cannot be embedded are called contorted. One natural candidate for the Hermitian metric could be the average of on [293] : , which reduces to on non-contorted 2-surfaces. Interestingly enough, can also be re-expressed by the integral ( 52 ) of the Nester–Witten 2-form [340] . Unfortunately, however, neither this metric nor the other suggestions appearing in the literature are conformally invariant. Thus, for contorted 2-surfaces, the definition of the quasi-local mass as the norm of the kinematical twistor (cf. ( 49 )) is ambiguous unless a natural is found. If is non-contorted, then the scalar product defines the totally anti-symmetric twistor , and for the four independent 2-surface twistors ,. . . , the contraction , and hence by ( 46 ) the determinant , is constant on . Nevertheless, can be constant even for contorted 2-surfaces for which is not. Thus, the totally anti-symmetric twistor can exist even for certain contorted 2-surfaces. Therefore, an alternative definition of the quasi-local mass might be based on ( 50 ) [354] . However, although the two mass definitions are equivalent in the linearized theory, they are different invariants of the kinematical twistor even in de Sitter or anti-de-Sitter spacetimes. Thus, if needed, the former notion of mass will be called the norm-mass, the latter the determinant-mass (denoted by ). If we want to have not only the notion of the mass but its reality is also expected, then we should ensure the Hermiticity of the kinematical twistor. But to formulate the Hermiticity condition ( 48 ), one also needs the infinity twistor. However, is not constant on even if it is non-contorted, thus in general it does not define any twistor on . One might take its average on (which can also be re-expressed by the integral of the Nester–Witten 2-form [340] ), but the resulting twistor would not be simple. In fact, even on 2-surfaces in de Sitter and anti-de Sitter spacetimes with cosmological constant the natural definition for is [299, 297, 354] , while on round spheres in spherically symmetric spacetimes it is [347] . Thus no natural simple infinity twistor has been found in curved spacetime. Indeed, Helfer claims that no such infinity twistor can exist [192] : Even if the spacetime is conformally flat (whenever the Hermitian metric exists) the Hermiticity condition would be fifteen algebraic equations for the (at most) twelve real components of the ‘would be’ infinity twistor. Then, since the possible kinematical twistors form an open set in the space of symmetric twistors, the Hermiticity condition cannot be satisfied even for non-simple s. However, in contrast to the linearized gravity case, the infinity twistor should not be given once and for all on some ‘universal’ twistor space, that may depend on the actual gravitational field. In fact, the 2-surface twistor space itself depends on the geometry of , and hence all the structures thereon also. Since in the Hermiticity condition ( 48 ) only the special combination of the infinity and metric twistors (the so-called ‘bar-hook’ combination) appears, it might still be hoped that an appropriate could be found for a class of 2-surfaces in a natural way [360] . However, as far as the present author is aware of, no real progress has been achieved in this way.7.2.4 The various limits
Obviously, the kinematical twistor vanishes in flat spacetime and, since the basic idea came from the linearized gravity, the construction gives the correct results in the weak field approximation. The nonrelativistic weak field approximation, i. e. the Newtonian limit, was clarified by Jeffryes [216] . He considers a one parameter family of spacetimes with perfect fluid source such that in the limit of the parameter one gets a Newtonian spacetime, and, in the same limit, the 2-surface lies in a hypersurface of the Newtonian time . In this limit the pointwise Hermitian scalar product is constant, and the norm-mass can be calculated. As could be expected, for the leading order term in the expansion of as a series of he obtained the conserved Newtonian mass. The Newtonian energy, including the kinetic and the Newtonian potential energy, appears as a order correction. The Penrose definition for the energy-momentum and angular momentum can be applied to the cuts of the future null infinity of an asymptotically flat spacetime [293, 299] . Then every element of the construction is built from conformally rescaled quantities of the non-physical spacetime. Since is shear-free, the 2-surface twistor equations on decouple, and hence the solution space admits a natural infinity twistor . It singles out precisely those solutions whose primary spinor parts span the asymptotic spin space of Bramson (see subsection 4.2.4 ), and they will be the generators of the energy-momentum. Although is contorted, and hence there is no natural Hermitian scalar product, there is a twistor with respect to which is Hermitian. Furthermore, the determinant is constant on , and hence it defines a volume 4-form on the 2-surface twistor space [360] . The energy-momentum coming from is just that of Bondi and Sachs. The angular momentum defined by is, however, new. It has a number of attractive properties. First, in contrast to definitions based on the Komar expression, it does not have the ‘factor-of-two anomaly’ between the angular momentum and the energy-momentum. Since its definition is based on the solutions of the 2-surface twistor equations (which can be interpreted as the spinor constituents of certain BMS vector fields generating boost-rotations) instead of the BMS vector fields themselves, it is free of supertranslation ambiguities. In fact, the 2-surface twistor space on reduces the BMS Lie algebra to one of its Poincare subalgebras. Thus the concept of the ‘translation of the origin’ is moved from null infinity to the twistor space (appearing in the form of a four parameter family of ambiguities in the potential for the shear ), and the angular momentum transforms just in the expected way under such a ‘translation of the origin’. As was shown in [125] , Penrose’s angular momentum can be considered as a supertranslation of previous definitions. The corresponding angular momentum flux through a portion of the null infinity between two cuts was calculated in [125, 191] and it was shown that this is precisely that given by Ashtekar and Streubel [28] (see also [321, 322, 124] ). The other way of determining the null infinity limit of the energy-momentum and angular momentum is to calculate them for the large spheres from the physical data, instead of the spheres at null infinity from the conformally rescaled data. These calculations were done by Shaw [323, 325] . At this point it should be noted that the limit of vanishes, and it is that yields the energy-momentum and angular momentum at infinity (see the remarks following eq. ( 15 )). The specific radiative solution for which the Penrose mass has been calculated is that of Robinson and Trautman [354] . The 2-surfaces for which the mass was calculated are the cuts of the geometrically distinguished outgoing null hypersurfaces . Tod found that, for given , the mass is independent of , as could be expected because of the lack of the incoming radiation. The large sphere limit of the 2-surface twistor space and the Penrose construction were investigated by Shaw in the Sommers [328] , the Ashtekar–Hansen [22] and the Beig–Schmidt [47] models of spatial infinity in [319, 320, 322] . Since no gravitational radiation is present near the spatial infinity, the large spheres are (asymptotically) non-contorted, and both the Hermitian scalar product and the infinity twistor are well defined. Thus the energy-momentum and angular momentum (and, in particular, the mass) can be calculated. In vacuum he recovered the Ashtekar–Hansen expression for the energy-momentum and angular momentum, and proved their conservation if the Weyl curvature is asymptotically purely electric. In the presence of matter the conservation of the angular momentum was investigated in [324] . The Penrose mass in asymptotically anti-de-Sitter spacetimes was studied by Kelly [228] . He calculated the kinematical twistor for spacelike cuts of the infinity , which is now a timelike 3-manifold in the non-physical spacetime. Since admits global 3-surface twistors (see the next subsection), is non-contorted. In addition to the Hermitian scalar product there is a natural infinity twistor, and the kinematical twistor satisfies the corresponding Hermiticity condition. The energy-momentum 4-vector coming from the Penrose definition is shown to coincide with that of Ashtekar and Magnon [26] . Therefore, the energy-momentum 4-vector is future pointing and timelike if there is a spacelike hypersurface extending to on which the dominant energy condition is satisfied. Consequently, . Kelly showed that is also non-negative and in vacuum it coincides with . In fact [360] , holds.7.2.5 The quasi-local mass of specific 2-surfaces
The Penrose mass has been calculated in a large number of specific situations. Round spheres are always non-contorted [358] , thus the norm-mass can be calculated. (In fact, axis-symmetric 2-surfaces in spacetimes with twist-free rotational Killing vector are non-contorted [217] .) The Penrose mass for round spheres reduces to the standard energy expression discussed in subsection 4.2.1 [354] . Thus every statement given in subsection 4.2.1 for round spheres is valid for the Penrose mass, and we do not repeat them. In particular, for round spheres in a slice of the Kantowski–Sachs spacetime this mass is independent of the 2-surfaces [351] . Interestingly enough, although these spheres cannot be shrunk to a point (thus the mass cannot be interpreted as ‘the 3-volume integral of some mass density’), the time derivative of the Penrose mass looks like the mass conservation equation: It is minus the pressure times the rate of change of the 3-volume of a sphere in flat space with the same area as [359] . In conformally flat spacetimes [354] the 2-surface twistors are just the global twistors restricted to , and the Hermitian scalar product is constant on . Thus the norm-mass is well defined. The construction works nicely even if global twistors exist only on a (say) spacelike hypersurface containing . These twistors are the so-called 3-surface twistors [354, 356] , which are solutions of certain (overdetermined) elliptic partial differential equations, the so-called 3-surface twistor equations, on . These equations are completely integrable (i. e. they admit the maximal number of linearly independent solutions, namely four) if and only if with its intrinsic metric and extrinsic curvature can be embedded, at least locally, into some conformally flat spacetime [356] . Such hypersurfaces are called non-contorted. It might be interesting to note that the non-contorted hypersurfaces can also be characterized as the critical points of the Chern–Simons functional built from the real Sen connection on the Lorentzian vector bundle or from the 3-surface twistor connection on the twistor bundle over [48, 345] . Returning to the quasi-local mass calculations, Tod showed that in vacuum the kinematical twistor for a 2-surface in a non-contorted depends only on the homology class of . In particular, if can be shrunk to a point then the corresponding kinematical twistor is vanishing. Since is non-contorted, is also non-contorted, and hence the norm–mass is well defined. This implies that the Penrose mass in the Schwarzschild solution is the Schwarzschild mass for any non-contorted 2-surface that can be deformed into a round sphere, and it is zero for those that do not link the black hole [358] . Thus, in particular, the Penrose mass can be zero even in curved spacetimes. A particularly interesting class of non-contorted hypersurfaces is that of the conformally flat time-symmetric initial data sets. Tod considered Wheeler’s solution of the time symmetric vacuum constraints describing ‘points at infinity’ (or, in other words, black holes) and 2-surfaces in such a hypersurface [354] . He found that the mass is zero if does not link any black hole, it is the mass of the -th black hole if links precisely the -th hole, it is if links precisely the -th and the -th holes, where is some appropriate measure of the distance of the holes, . . . , etc. Thus, the mass of the -th and -th holes as a single object is less than the sum of the individual masses, in complete agreement with our physical intuition that the potential energy of the composite system should contribute to the total energy with negative sign. Beig studied the general conformally flat time symmetric initial data sets describing ‘points at infinity’ [44] . He found a symmetric trace-free and divergence-free tensor field and, for any conformal Killing vector of the data set, defined the 2-surface flux integral of on . He showed that is conformally invariant, depends only on the homology class of , and, apart from numerical coefficients, for the ten (locally existing) conformal Killing vectors these are just the components of the kinematical twistor derived by Tod in [354] (and discussed in the previous paragraph). In particular, Penrose’s mass in Beig’s approach is proportional to the length of the ’s with respect to the Cartan-Killing metric of the conformal group of the hypersurface. Tod calculated the quasi-local mass for a large class of axis-symmetric 2-surfaces (cylinders) in various LRS Bianchi and Kantowski–Sachs cosmological models [359] and more general cylindrically symmetric spacetimes [361] . In all these cases the 2-surfaces are non-contorted, and the construction works. A technically interesting feature of these calculations is that the 2-surfaces have edges, i. e. they are not smooth submanifolds. The twistor equation is solved on the three smooth pieces of the cylinder separately, and the resulting spinor fields are required to be continuous at the edges. This matching reduces the number of linearly independent solutions to four. The projection parts of the resulting twistors, the s, are not continuous at the edges. It turns out that the cylinders can be classified invariantly to be hyperbolic, parabolic or elliptic. Then the structure of the quasi-local mass expressions is not simply ‘density’ ‘volume’, but they are proportional to a ‘type factor’ as well, where is the coordinate length of the cylinder. In the hyperbolic, parabolic and elliptic cases this factor is , 1 and , respectively, where is an invariant of the cylinder. The various types are interpreted as the presence of a positive, zero or negative potential energy. In the elliptic case the mass may be zero for finite cylinders. On the other hand, for static perfect fluid spacetimes (hyperbolic case) the quasi-local mass is positive. A particularly interesting spacetime is that describing cylindrical gravitational waves, whose presence is detected by the Penrose mass. In all these cases the determinant-mass has also been calculated and found to coincide with the norm-mass. A numerical investigation of the axis-symmetric Brill waves on the Schwarzschild background was presented in [67] . It was found that the quasi-local mass is positive, and it is very sensitive to the presence of the gravitational waves. Another interesting issue is the Penrose inequality for black holes (see subsection 13.2.1 ). Tod showed [357, 358] that for static black holes the Penrose inequality holds if the mass of the hole is defined to be the Penrose quasi-local mass of the spacelike cross section of the event horizon. The trick here is that is totally geodesic and conformal to the unit sphere, and hence it is non-contorted and the Penrose mass is well defined. Then the Penrose inequality will be a Sobolev type inequality for a non-negative function on the unit sphere. This inequality was tested numerically in [67] . Apart from the cuts of in radiative spacetimes, all the 2-surfaces discussed so far were non-contorted. The spacelike cross section of the event horizon of the Kerr black hole provides a contorted 2-surface [360] . Thus although the kinematical twistor can be calculated for this, the construction in its original form cannot yield any mass expression. The original construction has to be modified.7.2.6 Small surfaces
The properties of the Penrose construction that we have discussed are very remarkable and promising. However, the small surface calculations showed clearly some unwanted feature of the original construction [355, 229, 379] , and forced its modification. First, although the small spheres are contorted in general, the leading term of the pointwise Hermitian scalar product is constant: for any 2-surface twistors and [355, 229] . Since in non-vacuum spacetimes the kinematical twistor has only the ‘4-momentum part’ in the leading order with , the Penrose mass, calculated with the norm above, is just the expected mass in the leading order. Thus it is positive if the dominant energy condition is satisfied. On the other hand, in vacuum the structure of the kinematical twistor is(53) |
7.3 The modified constructions
Independently of the results of the small sphere calculations, Penrose claimed that in the Schwarzschild spacetime the quasi-local mass expression should yield the same zero value on 2-surfaces, contorted or not, which do not surround the black hole. (For the motivations and the arguments, see [295] .) Thus the original construction should be modified, and the negative results for the small spheres above strengthened this need. A much more detailed review of the various modifications is given by Tod in [360] .7.3.1 The ‘improved’ construction with the determinant
A careful analysis of the roots of the difficulties lead Penrose [295, 299] (see also [355, 229, 360] ) to suggest the modified definition for the kinematical twistor(54) |
7.3.2 Modification through Tod’s expression
These anomalies lead Penrose to modify slightly [296] . This modified form is based on Tod’s form of the kinematical twistor:(55) |
7.3.3 Mason’s suggestions
A beautiful property of the original construction was its connection with the Hamiltonian formulation of the theory [255] . Unfortunately, such a simple Hamiltonian interpretation is lacking for the modified constructions. Although the form of ( 55 ) is that of the integral of the Nester–Witten 2-form, and the spinor fields and could still be considered as the spinor constituents of the ‘quasi-Killing vectors’ of the 2-surface , their structure is not so simple because the factor itself depends on all of the four independent solutions of the 2-surface twistor equation in a rather complicated way. To have a simple Hamiltonian interpretation Mason suggested further modifications [255, 256] . He considers the four solutions , , of the 2-surface twistor equations, and uses these solutions in the integral ( 52 ) of the Nester–Witten 2-form. Since is a Hermitian bilinear form on the space of the spinor fields (see section 8 below), he obtains 16 real quantities as the components of the Hermitian matrix . However, it is not clear how the four ‘quasi-translations’ of should be found among the 16 vector fields (called ‘quasi-conformal Killing vectors’ of ) for which the corresponding quasi-local quantities could be considered as the quasi-local energy-momentum. Nevertheless, this suggestion leads us to the next class of quasi-local quantities.8 Approaches Based on the Nester–Witten 2-Form
We saw in subsection 3.2 that8.1 The Ludvigsen–Vickers construction
8.1.1 The definition
Suppose that the spacetime is asymptotically flat at future null infinity, and the closed spacelike 2-surface can be joined to future null infinity by a smooth null hypersurface . Let , the cut defined by the intersection of with the future null infinity. Then the null geodesic generators of define a smooth bijection between and the cut (and hence, in particular, ). We saw in subsection 4.2.4 that on the cut at the future null infinity we have the asymptotic spin space . The suggestion of Ludvigsen and Vickers [250] for the spin space on is to import the two independent solutions of the asymptotic twistor equations, i. e. the asymptotic spinors, from the future null infinity back to the 2-surface along the null geodesic generators of the null hypersurface . Their propagation equations, given both in terms of spinors and in the GHP formalism, are(56) |
(57) |
8.1.2 Remarks on the validity of the construction
Before discussing the usual questions about the properties of the construction (positivity, monotonicity, the various limits, etc.), we should make some general remarks. First, it is obvious that the Ludvigsen–Vickers energy-momentum in its form above cannot be defined in a spacetime which is not asymptotically flat at null infinity. Thus their construction is not genuinely quasi-local, because it depends not only on the (intrinsic and extrinsic) geometry of , but on the global structure of the spacetime as well. In addition, the requirement of the smoothness of the null hypersurface connecting the 2-surface to the null infinity is a very strong restriction. In fact, for general (even for convex) 2-surfaces in a general asymptotically flat spacetime conjugate points will develop along the (outgoing) null geodesics orthogonal to the 2-surface [290, 170] . Thus either the 2-surface must be near enough to the future null infinity (in the conformal picture), or the spacetime and the 2-surface must be nearly spherically symmetric (or the former cannot be ‘very much curved’ and the latter cannot be ‘very much bent’). This limitation yields that in general the original construction above does not have a small sphere limit. However, using the same propagation equations ( 56 )-( 57 ) one could define a quasi-local energy-momentum for small spheres [250, 64] . The basic idea is that there is a spin space at the vertex of the null cone in the spacetime whose spacelike cross section is the actual 2-surface, and the Ludvigsen–Vickers spinors on are defined by propagating these spinors from the vertex to via ( 56 )-( 57 ). This definition works in arbitrary spacetime, but the 2-surface cannot be extended to a large sphere near the null infinity, and it is still not genuinely quasi-local.8.1.3 Monotonicity, mass-positivity and the various limits
Once the Ludvigsen–Vickers spinors are given on a spacelike 2-surface of constant affine parameter in the outgoing null hypersurface , then they are uniquely determined on any other spacelike 2-surface in , too, i. e. the propagation law ( 56 )-( 57 ) defines a natural isomorphism between the space of the Ludvigsen–Vickers spinors on different 2-surfaces of constant affine parameter in the same . ( need not be a Bondi type coordinate.) This makes it possible to compare the components of the Ludvigsen–Vickers energy-momenta on different surfaces. In fact [250] , if the dominant energy condition is satisfied (at least on ), then for any Ludvigsen–Vickers spinor and affine parameter values one has , and the difference can be interpreted as the energy flux of the matter and the gravitational radiation through between and . Thus both and are increasing with (‘mass-gain’). A similar monotonicity property (‘mass-loss’) can be proven on ingoing null hypersurfaces, but then the propagation law ( 56 )-( 57 ) should be replaced by and . Using these equations the positivity of the Ludvigsen–Vickers mass was proven in various special cases in [250] . Concerning the positivity properties of the Ludvigsen–Vickers mass and energy, first it is obvious by the remarks on the nature of the propagation law ( 56 )-( 57 ) that in Minkowski spacetime the Ludvigsen–Vickers energy-momentum is vanishing. However, in the proof of the non-negativity of the Dougan–Mason energy (discussed in subsection 8.2 ) only the part of the propagation equations is used. Therefore, as realized by Bergqvist [59] , the Ludvigsen–Vickers energy-momenta (both based on the asymptotic and the point spinors) are also future directed and nonspacelike if is the boundary of some compact spacelike hypersurface on which the dominant energy condition is satisfied and is weakly future convex (or at least ). Similarly, the Ludvigsen–Vickers definitions share the rigidity properties proven for the Dougan–Mason energy-momentum [338] : Under the same conditions the vanishing of the energy-momentum implies the flatness of the domain of dependence of . In the weak field approximation [250] the difference is just the integral of on the portion of between the two 2-surfaces, where is the linearized energy-momentum tensor: The increment of on is due only to the flux of the matter energy-momentum. Since the Bondi–Sachs energy-momentum can be written as the integral of the Nester–Witten 2-form on the cut in question at the null infinity with the asymptotic spinors, it is natural to expect that the first version of the Ludvigsen–Vickers energy-momentum tends to that of Bondi and Sachs. It was shown in [250, 325] that this expectation is, in fact, correct. The Ludvigsen–Vickers mass was calculated for large spheres both for radiative and stationary spacetimes with and accuracy, respectively, in [323, 325] . Finally, on a small sphere of radius in non-vacuum the second definition gives [64] the expected result ( 28 ), while in vacuum [64, 344] it is(58) |
8.2 The Dougan–Mason constructions
8.2.1 Holomorphic/anti-holomorphic spinor fields
The original construction of Dougan and Mason [123] was introduced on the basis of sheaf-theoretical arguments. Here we follow a slightly different, more ‘pedestrian’ approach, based mostly on [338, 340] . Following Dougan and Mason we define the spinor field to be anti-holomorphic in case , or holomorphic if . Thus, this notion of holomorphicity/anti-holomorphicity is referring to the connection on . While the notion of the holomorphicity/anti-holomorphicity of a function on does not depend on whether the or the operator is used, for tensor or spinor fields it does. Although the vectors and are not uniquely determined (because their phase is not fixed), the notion of the holomorphicity/anti-holomorphicity is well defined, because the defining equations are homogeneous in and . Next suppose that there are at least two independent solutions of . If and are any two such solutions, then , and hence by Liouville’s theorem is constant on . If this constant is not zero, then we call generic, if it is zero then will be called exceptional. Obviously, holomorphic on a generic cannot have any zero, and any two holomorphic spinor fields, e. g. and , span the spin space at each point of (and they can be chosen to form a normalized spinor dyad with respect to on the whole of ). Expanding any holomorphic spinor field in this frame, the expanding coefficients turn out to be holomorphic functions, and hence constant. Therefore, on generic 2-surfaces there are precisely two independent holomorphic spinor fields. In the GHP formalism the condition of the holomorphicity of the spinor field is that its components be in the kernel of . Thus for generic 2-surfaces with the constant would be a natural candidate for the spin space above. For exceptional 2-surfaces the kernel space is either two dimensional but does not inherit a natural spin space structure, or it is higher than two dimensional. Similarly, the symplectic inner product of any two anti-holomorphic spinor fields is also constant, one can define generic and exceptional 2-surfaces as well, and on generic surfaces there are precisely two anti-holomorphic spinor fields. The condition of the anti-holomorphicity of is . Then could also be a natural choice. Note that since the spinor fields whose holomorphicity/anti-holomorphicity is defined are unprimed, and these correspond to the anti-holomorphicity/holomorphicity, respectively, of the primed spinor fields of Dougan and Mason. Thus the main question is whether there exist generic 2-surfaces, and if they do, whether they are ‘really generic’, i. e. whether most of the physically important surfaces are generic or not.8.2.2 The genericity of the generic 2-surfaces
are first order elliptic differential operators on certain vector bundles over the compact 2-surface , and their index can be calculated: , where is the genus of . Therefore, for there are at least two linearly independent holomorphic and at least two linearly independent anti-holomorphic spinor fields. The existence of the holomorphic/anti-holomorphic spinor fields on higher genus 2-surfaces is not guaranteed by the index theorem. Similarly, the index theorem does not guarantee that is generic either: If the geometry of is very special then the two holomorphic/anti-holomorphic spinor fields (which are independent as solutions of ) might be proportional to each other. For example, future marginally trapped surfaces (i. e. for which ) are exceptional from the point of view of holomorphic, and past marginally trapped surfaces ( ) from the point of view of anti-holomorphic spinors. Furthermore, there are surfaces with at least three linearly independent holomorphic/anti-holomorphic spinor fields. However, small generic perturbations of the geometry of an exceptional 2-surface with topology make generic. Finally, we note that several first order differential operators can be constructed from the chiral irreducible parts and of , given explicitly by ( 25 ). However, only four of them, the Dirac–Witten operator , the twistor operator , and the holomorphy and anti-holomorphy operators , are elliptic (which ellipticity, together with the compactness of , would guarantee the finiteness of the dimension of their kernel), and it is only that have two-complex dimensional kernel in the generic case. This purely mathematical result gives some justification for the choices of Dougan and Mason: The spinor fields that should be used in the Nester–Witten 2-form are either holomorphic or anti-holomorphic. The construction does not work for exceptional 2-surfaces.8.2.3 Positivity properties
One of the most important properties of the Dougan–Mason energy-momenta is that they are future pointing nonspacelike vectors, i. e. the corresponding masses and energies are non-negative. Explicitly [123] , if is the boundary of some compact spacelike hypersurface on which the dominant energy condition holds, furthermore if is weakly future convex (in fact, is enough), then the holomorphic Dougan–Mason energy-momentum is a future pointing non-spacelike vector, and, analogously, the anti-holomorphic energy-momentum is future pointing and non-spacelike if . As Bergqvist [59] stressed (and we noted in subsection 8.1.3 ), Dougan and Mason used only the (and in the anti-holomorphic construction the ) half of the ‘propagation law’ in their positivity proof. The other half is needed only to ensure the existence of two spinor fields. Thus that might be ( 56 ) of the Ludvigsen–Vickers construction, or in the holomorphic Dougan–Mason construction, or even for some constant , a ‘deformation’ of the holomorphicity considered by Bergqvist [59] . In fact, the propagation law may even be for any spinor field satisfying . This ensures the positivity of the energy under the same conditions and that is still constant on for any two solutions and , making it possible to define the norm of the resulting energy-momentum, i. e. the mass. In the asymptotically flat spacetimes the positive energy theorems have a rigidity part too, namely the vanishing of the energy-momentum (and, in fact, even the vanishing of the mass) implies flatness. There are analogous theorems for the Dougan–Mason energy-momenta too [338, 340] . Namely, under the conditions of the positivity proof8.2.4 The various limits
Both definitions give the same standard expression for round spheres [122] . Although the limit of the Dougan–Mason masses for round spheres in Reissner–Nordstrom spacetime gives the correct irreducible mass of the Reissner–Nordstrom black hole on the horizon, the constructions do not work on the surface of bifurcation itself, because that is an exceptional 2-surface. Unfortunately, without additional restrictions (e. g. the spherical symmetry of the 2-surfaces in a spherically symmetric spacetime) the mass of the exceptional 2-surfaces cannot be defined in a limiting process, because, in general, the limit depends on the family of generic 2-surfaces approaching the exceptional one [340] . Both definitions give the same, expected results in the weak field approximation and for large spheres at spatial infinity: Both tend to the ADM energy-momentum [123] . In non-vacuum both definitions give the same, expected expression ( 28 ) for small spheres, in vacuum they coincide in the order with that of Ludvigsen and Vickers, but in the order they differ from each other: The holomorphic definition gives ( 58 ), but in the analogous expression for the anti-holomorphic energy-momentum the numerical coefficient is replaced by [122] . The Dougan–Mason energy-momenta have also been calculated for large spheres of constant Bondi type radial coordinate value near future null infinity [122] . While the anti-holomorphic construction tends to the Bondi–Sachs energy-momentum, the holomorphic one diverges in general. In stationary spacetimes they coincide and both give the Bondi–Sachs energy-momentum. At the past null infinity it is the holomorphic construction which reproduces the Bondi–Sachs energy-momentum and the anti-holomorphic diverges. We close this subsection with some caution and general comments on a potential gauge ambiguity in the calculation of the various limits. By the definition of the holomorphic and anti-holomorphic spinor fields they are associated with the 2-surface only. Thus if is another 2-surface, then there is no natural isomorphism between the space – for example of the anti-holomorphic spinor fields on – and on , even if both surfaces are generic and hence there are isomorphisms between them. This (apparently ‘only theoretical’) fact has serious pragmatic consequences. In particular, in the small or large sphere calculations we compare the energy-momenta, and hence the holomorphic or anti-holomorphic spinor fields also, on different surfaces. For example [344] , in the small sphere approximation every spin coefficient and spinor component in the GHP dyad and metric component in some fixed coordinate system is expanded as a series of , like . Substituting all such expansions and the asymptotic solutions of the Bianchi identities for the spin coefficients and metric functions into the differential equations defining the holomorphic/anti-holomorphic spinors, we obtain a hierarchical system of differential equations for the expansion coefficients , , . . . , etc. It turns out that the solutions of this system of equations with accuracy form a rather than the expected two complex dimensional space. of these solutions are ‘gauge’ solutions, and they correspond in the approximation with given accuracy to the unspecified isomorphism between the space of the holomorphic/anti-holomorphic spinor fields on surfaces of different radii. Obviously, similar ‘gauge’ solutions appear in the large sphere expansions, too. Therefore, without additional gauge fixing, in the expansion of a quasi-local quantity only the leading non-trivial term will be gauge-independent. In particular, the order correction in ( 58 ) for the Dougan–Mason energy-momenta is well defined only as a consequence of a natural gauge choice. Similarly, the higher order corrections in the large sphere limit of the anti-holomorphic Dougan–Mason energy-momentum are also ambiguous unless a ‘natural’ gauge choice is made. Such a choice is possible in stationary spacetimes.Recall that, similarly, we did not have any natural isomorphism between the 2-surface twistor spaces, discussed in subsection 7.2.1 , on different 2-surfaces.
Clearly, for the Ludvigsen–Vickers energy-momentum no such ambiguity is present, because the part ( 56 ) of their propagation law defines a natural isomorphism between the space of the Ludvigsen–Vickers spinors on the different 2-surfaces.
8.3 A specific construction for the Kerr spacetime
Logically, this specific construction perhaps would have to be presented only in section 12 , but the technique that it is based on may justify its placing here. By investigating the propagation law ( 56 )-( 57 ) of Ludvigsen and Vickers, for the Kerr spacetimes Bergqvist and Ludvigsen constructed a natural flat, (but non-symmetric) metric connection [65] . Writing the new covariant derivative in the form , the ‘correction’ term could be given explicitly in terms of the GHP spinor dyad (adapted to the two principal null directions), the spin coefficients , and , and the curvature component . admits a potential [66] : , where . However, this potential has the structure appearing in the form of the metric for the Kerr–Schild spacetimes, where is the flat metric. In fact, the flat connection above could be introduced for general Kerr–Schild metrics [165] , and the corresponding ‘correction term’ could be used to find easily the Lanczos potential for the Weyl curvature [9] . Since the connection is flat and annihilates the spinor metric , there are precisely two linearly independent spinor fields, say and , that are constant with respect to and form a normalized spinor dyad. These spinor fields are asymptotically constant. Thus it is natural to choose the spin space to be the space of the -constant spinor fields, independently of the 2-surface . A remarkable property of these spinor fields is that the Nester–Witten 2-form built from them is closed: . This implies that the quasi-local energy-momentum depends only on the homology class of , i. e. if and are 2-surfaces such that they form the boundary of some hypersurface in , then , and if is the boundary of some hypersurface, then . In particular, for two-spheres that can be shrunk to a point the energy-momentum is zero, but for those that can be deformed to a cut of the future null infinity the energy-momentum is that of Bondi and Sachs.9 Quasi-Local Spin-Angular Momentum
In this section we review three specific quasi-local spin-angular momentum constructions that are (more or less) ‘quasi-localizations’ of Bramson’s expression at null infinity. Thus the quasi-local spin-angular momentum for the closed, orientable spacelike 2-surface will be sought in the form ( 17 ). Before considering the specific constructions themselves we summarize the most important properties of the general expression of ( 17 ). Since the most detailed discussion of ( 17 ) is given probably in [344, 347] , the subsequent discussions will be based on them. First, observe that the integral depends on the spinor dyad algebraically, thus it is enough to specify the dyad only at the points of . Obviously, transforms like a symmetric second rank spinor under constant transformations of the dyad . Second, suppose that the spacetime is flat, and let be constant. Then the corresponding 1-form basis is the constant Cartesian one, which consists of exact 1-forms. Then since the Bramson superpotential is the anti-self-dual part (in the name indices) of , which is also exact, for such spinor bases ( 17 ) gives zero. Therefore, the integral of Bramson’s superpotential ( 17 ) measures the non-integrability of the 1-form basis , i. e. is a measure of how much the actual 1-form basis is ‘distorted’ by the curvature relative to the constant basis of Minkowski spacetime. Thus the only question is how to specify a spin frame on to be able to interpret as angular momentum. It seems natural to choose those spinor fields that were used in the definition of the quasi-local energy-momenta in the previous section. At first sight this may appear to be only an ad hoc idea, but, recalling that in section 8 we interpreted the elements of the spin spaces as the ‘spinor constituents of the quasi-translations of ’, we can justify such a choice. Based on our experience with the superpotentials for the various conserved quantities, the quasi-local angular momentum can be expected to be the integral of something like ‘superpotential’ ‘quasi-rotation generator’, and the ‘superpotential’ is some expression in the first derivative of the basic variables, actually the tetrad or spinor basis. Since, however, Bramson’s superpotential is an algebraic expression of the basic variables, and the number of the derivatives in the expression for the angular momentum should be one, the angular momentum expressions based on Bramson’s superpotential must contain the derivative of the ‘quasi-rotations’, i. e. (possibly a combination of) the ‘quasi-translations’. Since, however, such an expression cannot be sensitive to the ‘change of the origin’, they can be expected to yield only the spin part of the angular momentum. The following two specific constructions differ from each other only in the choice for the spin space , and correspond to the energy-momentum constructions of the previous section. The third construction (valid only in the Kerr spacetimes) is based on the sum of two terms, where one is Bramson’s expression, and uses the spinor fields of subsection 8.3 . Thus the present section is not independent of section 8 , and for the discussion of the choice of the spin spaces we refer to that. Another suggestion for the quasi-local spatial angular momentum, proposed by Liu and Yau [245] , will be introduced in subsection 10.4.1 .9.1 The Ludvigsen–Vickers angular momentum
Under the conditions that ensured the Ludvigsen–Vickers construction for the energy-momentum would work in subsection 8.1 , the definition of their angular momentum is straightforward [250] . Since in Minkowski spacetime the Ludvigsen–Vickers spinors are just the restriction to of the constant spinor fields, by the general remark above the Ludvigsen–Vickers spin-angular momentum is zero in Minkowski spacetime. Using the asymptotic solution of the Einstein–Maxwell equations in a Bondi type coordinate system it has been shown in [250] that the Ludvigsen–Vickers spin-angular momentum tends to that of Bramson at future null infinity. For small spheres [344] in non-vacuum it reproduces precisely the expected result ( 29 ), and in vacuum it is(59) |
9.2 Holomorphic/anti-holomorphic spin-angular momenta
Obviously, the spin-angular momentum expressions based on the holomorphic and anti-holomorphic spinor fields [342] on generic 2-surfaces are genuinely quasi-local. Since in Minkowski spacetime the restriction of the two constant spinor fields to any 2-surface are constant, and hence holomorphic and anti-holomorphic at the same time, both the holomorphic and anti-holomorphic spin-angular momenta are vanishing. Similarly, for round spheres both definitions give zero [347] , as it could be expected in a spherically symmetric system. The anti-holomorphic spin-angular momentum has already been calculated for axis-symmetric 2-surfaces for which the anti-holomorphic Dougan–Mason energy-momentum is null, i. e. for which the corresponding quasi-local mass is zero. (As we saw in subsection 8.2.3 , this corresponds to a pp-wave geometry and pure radiative matter fields on [338, 340] .) This null energy-momentum vector turned out to be an eigenvector of the anti-symmetric spin-angular momentum tensor , which, together with the vanishing of the quasi-local mass, is equivalent to the proportionality of the (null) energy-momentum vector and the Pauli–Lubanski spin [342] , where the latter is defined by(60) |
9.3 A specific construction for the Kerr spacetime
The angular momentum of Bergqvist and Ludvigsen [66] for the Kerr spacetime is based on their special flat, non-symmetric but metric connection explained briefly in subsection 8.3 , but their idea is not simply the use of the two -constant spinor fields in Bramson’s superpotential. Rather, in the background of their approach there are twistor-theoretical ideas. (The twistor-theoretic aspects of the analogous flat connection for the general Kerr–Schild class are discussed in [165] .) The main idea is that while the energy-momentum is a single four-vector in the dual of the Hermitian subspace of , the angular momentum is not only an anti-symmetric tensor over the same space, but should depend on the ‘origin’, a point in a four dimensional affine space as well, and should transform in a specific way under the translation of the ‘origin’. Bergqvist and Ludvigsen defined the affine space to be the space of the solutions of , and showed that is, in fact, a real, four dimensional affine space. Then, for a given , to each -constant spinor field they associate a primed spinor field by . This turns out to satisfy the modified valence 1 twistor equation . Finally, they form the 2-form(61) |
10 The Hamilton–Jacobi Method
If one is concentrating only on the introduction and study of the properties of the quasi-local quantities, but not interested in the detailed structure of the quasi-local (Hamiltonian) phase space, then perhaps the most natural way to derive the general formulae is to follow the Hamilton–Jacobi method. This was done by Brown and York in deriving their quasi-local energy expression [93, 94] . However, the Hamilton–Jacobi method in itself does not yield any specific construction. Rather, the resulting general expression is similar to a superpotential in the Lagrangian approaches, which should be completed by a choice for the reference configuration and for the generator vector field of the physical quantity (see subsection 3.3.3 ). In fact, the ‘Brown–York quasi-local energy’ is not a single expression with a single well defined prescription for the reference configuration. The same general formula with several other, mathematically inequivalent definitions for the reference configurations are still called the ‘Brown–York energy’. A slightly different general expression was used by Kijowski [231] , Epp [129] and Liu and Yau [245] . Although the former follows a different route to derive his expression and the latter two are not connected directly to the canonical analysis (and, in particular, to the Hamilton–Jacobi method), the formalism and techniques that are used justify their presentation in this section. The present section is based mostly on the original papers [93, 94] by Brown and York. Since, however, this is the most popular approach to finding quasi-local quantities and is the subject of very active investigations, especially from the point of view of the applications in black hole physics, this section is perhaps less complete than the previous ones. The expressions of Kijowski, Epp and Liu and Yau will be treated in the formalism of Brown and York.10.1 The Brown–York expression
10.1.1 The main idea
To motivate the main idea behind the Brown–York definition [93, 94] , let us consider first a classical mechanical system of degrees of freedom with configuration manifold and Lagrangian (i. e. the Lagrangian is assumed to be first order and may depend on time explicitly). For given initial and final configurations, and , respectively, the corresponding action functional is , where is a smooth curve in from to with tangent at . (The pair may be called a history or world line in the ‘spacetime’ .) Let be a smooth 1 parameter deformation of this history, i. e. for which , and for some . Then, denoting the derivative with respect to the deformation parameter at by , one has the well known expression(62) |
10.1.2 The variation of the action and the surface stress-energy tensor
The main idea of Brown and York [93, 94] is to calculate the analogous variation of an appropriate first order action of general relativity (or of the coupled matter+gravity system) and isolate the boundary term that could be analogous to the energy above. To formulate this idea mathematically, they considered a compact spacetime domain with topology such that correspond to compact spacelike hypersurfaces ; these form a smooth foliation of and the 2-surfaces (corresponding to ) form a foliation of the timelike 3-boundary of . Note that this is not a globally hyperbolic domain. To ensure the compatibility of the dynamics with this boundary, the shift vector is usually chosen to be tangent to on . The orientation of is chosen to be outward pointing, while the normals both of and to be future pointing. The metric and extrinsic curvature on will be denoted, respectively, by and , those on by and . The primary requirement of Brown and York on the action is to provide a well defined variational principle for the Einstein theory. This claim leads them to choose for the ‘trace K action’ (or, in the present notation, rather the ‘trace action’) for general relativity [385, 386, 369] , and the action for the matter fields may be included. (For the minimal, non-derivative couplings the presence of the matter fields does not alter the subsequent expressions.) However, as Geoff Hayward pointed out [173] , to have a well defined variational principle, the ‘trace action’ should in fact be completed by two 2-surface integrals, one on and the other on . Otherwise, as a consequence of the edges and , called the ‘joints’ (i. e. the non-smooth parts of the boundary ), the variation of the metric at the points of the edges and could not be arbitrary. (See also [172, 231, 75, 92] , where the ‘orthogonal boundaries assumption’ is also relaxed.) Let and be the scalar product of the outward pointing normal of and the future pointing normal of and of , respectively. Then, varying the spacetime metric, for the variation of the corresponding principal function they obtained(63) |
(64) |
(65) |
In the original papers Brown and York assumed that the leaves of the foliation of were orthogonal to (‘orthogonal boundaries assumption’).
10.1.3 The general form of the Brown–York quasi-local energy
The 3+1 decomposition of the spacetime metric yields a 2+1 decomposition of the metric , too. Let and be the lapse and the shift of this decomposition on . Then the corresponding decomposition of defines the energy, momentum and spatial stress surface densities according to(66) |
(67) |
(68) |
(69) |
(70) |
(71) |
10.1.4 Further properties of the general expressions
As we noted, , and depend on the boost-gauge that the timelike boundary defines on . Lau clarified how these quantities change under a boost gauge transformation, where the new boost-gauge is defined by the timelike boundary of another domain such that the particular 2-surface is a leaf of the foliation of too [240] : If is another foliation of such that and is orthogonal to , then the new , and are built from the old , and and the 2+1 pieces on of the canonical momentum , defined on . Apart from the contribution of , these latter quantities are(72) |
(73) |
(74) |
The paper gives a clear, well readable summary of these earlier results.
10.1.5 The Hamiltonians
If we can write the action of our mechanical system into the canonical form , then it is straightforward to read off the Hamiltonian of the system. Thus, having accepted the trace action as the action for general relativity, it is natural to derive the corresponding Hamiltonian in the analogous way. Following this route Brown and York derived the Hamiltonian, corresponding to the ‘basic’ (or non-referenced) action too [94] . They obtained the familiar integral of the sum of the Hamiltonian and the momentum constraints, weighted by the lapse and the shift , respectively, plus , given by ( 69 ), as a boundary term. This result is in complete agreement with the expectations, as their general quasi-local quantities can also be recovered as the value of the Hamiltonian on the constraint surface. (See also [75] .) This Hamiltonian was investigated further in [92] . Here all the boundary terms that appear in the variation of their Hamiltonian are determined and decomposed with respect to the 2-surface . It is shown that the change of the Hamiltonian under a boost of yields precisely the boosts of the energy and momentum surface density discussed above. Hawking, Horowitz and Hunter also derived the Hamiltonian from the trace action both with the orthogonal [171] and non-orthogonal boundaries assumptions [172] . They allowed matter fields , whose dynamics is governed by a first order action , to be present. However, they treated the reference configuration in a different way. In the traditional canonical analysis of the fields and the geometry based on a non-compact (for example in the asymptotically flat case) one has to impose certain fall-off conditions that ensure the finiteness of the action, the Hamiltonian, etc. This finiteness requirement excludes several potentially interesting field+gravity configurations from our investigations. In fact, in the asymptotically flat case we compare the actual matter+gravity configurations with the flat spacetime+vanishing matter fields configuration. Hawking and Horowitz generalized this picture by choosing a static, but otherwise arbitrary solution , of the field equations, considered the timelike boundary of to be a timelike cylinder ‘near the infinity’, and considered the action and those matter+gravity configurations which induce the same value on as and . Its limit as is ‘pushed out to infinity’ can be finite even if the limit of the original (i. e. non-referenced) action is infinite. Although in the non-orthogonal boundaries case the Hamiltonian derived from the non-referenced action contains terms coming from the ‘joints’, by the boundary conditions at they are canceled from the referenced Hamiltonian. This latter Hamiltonian coincides with that obtained in the orthogonal boundaries case. Both the ADM and the Abbott–Deser energy can be recovered from this Hamiltonian [171] , and the quasi-local energy for spheres in domains with non-orthogonal boundaries in the Schwarzschild solution is also calculated [172] . A similar Hamiltonian, including the ‘joints’ or ‘corner’ terms, was obtained by Francaviglia and Raiteri [137] for the vacuum Einstein theory (and for Einstein–Maxwell systems in [4] ), using a Noether charge approach. Their formalism, using the language of jet bundles, is, however, slightly more sophisticated than that common in general relativity. Booth and Fairhurst [71] reexamined the general form of the Brown–York energy and angular momentum from a Hamiltonian point of view. Their starting point is the observation that the domain is not isolated from its environment, thus the quasi-local Hamiltonian cannot be time independent. Therefore, instead of the standard Hamiltonian formalism for the autonomous systems, a more general formalism, based on the extended phase space, must be used. This phase space consists of the usual bulk configuration and momentum variables on the typical 3-manifold and the time coordinate , the space coordinates on the 2-boundary and their conjugate momenta and , respectively. Their second important observation is that the Brown–York boundary conditions are too restrictive: The 2-metric, the lapse and the shift need not to be fixed but their variations corresponding to diffeomorphisms on the boundary must be allowed. Otherwise diffeomorphisms that are not isometries of the 3-metric on cannot be generated by any Hamiltonian. Relaxing the boundary conditions appropriately, they show that there is a Hamiltonian on the extended phase space which generates the correct equations of motions, and the quasi-local energy and angular momentum expression of Brown and York are just (minus) the momentum conjugate to the time coordinate . The only difference between the present and the original Brown–York expressions is the freedom in the functional form of the unspecified reference term: Because of the more restrictive boundary conditions of Brown and York their reference term is less restricted. Choosing the same boundary conditions in both approaches the resulting expressions coincide completely.Thus, in principle, we would have to report on their investigations in the next section. Nevertheless, since essentially they re-derive and justify the results of Brown and York following only a different route, we discuss their results here.
10.1.6 The flat space and light cone references
The quasi-local quantities introduced above become well defined only if the subtraction term in the principal function is specified. The usual interpretation of a choice for is the calibration of the quasi-local quantities, i. e. fixing where to take their zero value. The only restriction on that we had is that it must be a functional of the metric on the timelike boundary . To specify , it seems natural to expect that the principal function be zero in Minkowski spacetime [154, 93] . Then would be the integral of the trace of the extrinsic curvature of if it were embedded in Minkowski spacetime with the given intrinsic metric . However, a general Lorentzian 3-manifold cannot be isometrically embedded, even locally, into the Minkowski spacetime. (For a detailed discussion of this embeddability see [93] and subsection 10.1.8 .) Another assumption on might be the requirement of the vanishing of the quasi-local quantities, or of the energy and momentum surface densities, or only of the energy surface density , in some reference spacetime, e. g. in Minkowski or in anti-de-Sitter spacetime. Assuming that depends on the lapse and shift linearly, the functional derivatives and depend only on the 2-metric and on the boost-gauge that defined on . Therefore, and take the form ( 71 ), and by the requirement of the vanishing of in the reference spacetime it follows that should be the trace of the extrinsic curvature of in the reference spacetime. Thus it would be natural to fix as the trace of the extrinsic curvature of when is embedded isometrically into the reference spacetime. However, this embedding is far from being unique (since, in particular, there are two independent normals of in the spacetime and it would not be fixed which normal should be used to calculate ), and hence the construction would be ambiguous. On the other hand, one could require to be embedded into flat Euclidean 3-space, i. e. into a spacelike hyperplane of Minkowski spacetime. This is the choice of Brown and York [93, 94] . In fact, at least for a large class of 2-surfaces , such an embedding exists and is unique: If and the metric is and has everywhere positive scalar curvature, then there is an isometric embedding of into the flat Euclidean 3-space [190] , and apart from rigid motions this embedding is unique [330] . The requirement that the scalar curvature of the 2-surface must be positive can be interpreted as some form of the convexity, as in the theory of surfaces in the Euclidean space. However, there are counterexamples even to local isometric embeddability when this convexity condition is violated [265] . A particularly interesting 2-surface that cannot be isometrically embedded into the flat 3-space is the event horizon of the Kerr black hole if the angular momentum parameter exceeds the irreducible mass (but is still not greater than the mass parameter ), i. e. if [327] . Thus, the construction works for a large class of 2-surfaces, but certainly not for every potentially interesting 2-surface. The convexity condition is essential. It is known that the (local) isometric embeddability of into flat 3-space with extrinsic curvature is equivalent to the Gauss–Codazzi–Mainardi equations and . Here is the intrinsic Levi-Civita covariant derivative and is the corresponding curvature scalar on determined by . Thus, for given and (actually the flat) embedding geometry, these are three equations for the three components of , and hence, if the embedding exists, determines . Therefore, the subtraction term can also be interpreted as a solution of an under-determined elliptic system which is constrained by a nonlinear algebraic equation. In this form the definition of the reference term is technically analogous to the definition of those in sections 7 , 8 and 9 , but, by the non-linearity of the equations, in practice it is much more difficult to find the reference term than the spinor fields in the constructions of sections 7 , 8 and 9 . Accepting this choice for the reference configuration, the reference gauge potential will be zero in the boost-gauge in which the timelike normal of in the reference Minkowski spacetime is orthogonal to the spacelike 3-plane, because this normal is constant. Thus, to summarize, for convex 2-surfaces the flat space reference of Brown and York is uniquely determined, is determined by this embedding, and . Then , from which can be calculated (if needed). The procedure is similar if, instead of a spacelike hyperplane of Minkowski spacetime, a spacelike hypersurface of constant curvature (for example in the de-Sitter or anti-de-Sitter spacetime) is used. The only difference is that extra (known) terms appear in the Gauss–Codazzi–Mainardi equations. Brown, Lau and York considered another prescription for the reference configuration as well [91, 241, 242] . In this approach the 2-surface is embedded into the light cone of a point of the Minkowski or anti-de Sitter spacetime instead of a spacelike hypersurface of constant curvature. The essential difference between the new (‘light cone reference’) and the previous (‘flat space reference’) prescriptions is that the embedding into the light cone is not unique, but the reference term may be given explicitly, in a closed form. The positivity of the Gauss curvature of the intrinsic geometry of is not needed. In fact, by a result of Brinkmann [88] , every locally conformally flat Riemannian -geometry is locally isometric to an appropriate cut of a light cone of the dimensional Minkowski spacetime (see also [129] ). To achieve uniqueness some extra condition must be imposed. This may be the requirement of the vanishing of the ‘normal momentum density’ in the reference spacetime [241, 242] , yielding , where is the Ricci scalar of and is the cosmological constant of the reference spacetime. The condition defines something like a ‘rest frame’ in the reference spacetime. Another, considerably more complicated choice for the light cone reference term is used in [91] .The problem to characterize this embeddability is known as the Weyl problem of differential geometry.
10.1.7 Further properties and the various limits
Although the general, non-referenced expressions are additive, the prescription for the reference term destroys the additivity in general. In fact, if and are 2-surfaces such that is connected and two dimensional (more precisely, it has a non-empty open interior for example in ), then in general (overline means topological closure) is not guaranteed to be embeddable into the flat 3-space, and even if it is embeddable then the resulting reference term differs from the reference terms and determined from the individual embeddings. As it is noted in [75] , the Brown–York energy with the flat space reference configuration is not zero in Minkowski spacetime in general. In fact, in the standard spherical polar coordinates let be the spacelike hyperboloid , the hyperplane and , the sphere of radius in the hyperplane. Then the trace of the extrinsic curvature of in and in is and , respectively. Therefore, the Brown–York quasi-local energy (with the flat 3-space reference) associated with and the normals of on is . Similarly, the Brown–York quasi-local energy with the light cone references in [241] and in [91] is also negative for such surfaces with the boosted observers [?] . Recently, Shi and Tam [326] proved interesting theorems in Riemannian 3-geometries, which can be used to prove positivity of the Brown–York energy if the 2-surface is a boundary of some time symmetric spacelike hypersurface on which the dominant energy condition holds. In the time symmetric case this energy condition is just the condition that the scalar curvature be non-negative. The key theorem of Shi and Tam is the following: Let be a compact, smooth Riemannian 3-manifold with non-negative scalar curvature and smooth 2-boundary such that each connected component, , of is homeomorphic to and the scalar curvature of the induced 2-metric on is strictly positive. Then for each component holds, where is the trace of the extrinsic curvature of in with respect to the outward directed normal, and is the trace of the extrinsic curvature of in the flat Euclidean 3-space when is isometrically embedded. Furthermore, if in these inequalities the equality holds for at least one , then itself is connected and is flat. The energy expression for round spheres in spherically symmetric spacetimes was calculated in [94, 75] . In the spherically symmetric metric discussed in subsection 4.2.1 on the round spheres the Brown–York energy with the flat space reference and fleet of observers on is . In particular, it is for the Schwarzschild solution. This deviates from the standard round sphere expression, and, for the horizon of the Schwarzschild black hole it is (instead of the expected ). (The energy has also been calculated explicitly for boosted foliations of the Schwarzschild solution and for round spheres in isotropic cosmological models [92] .) The Newtonian limit can be derived from this by assuming that is the mass of a fluid ball of radius and is small: It is . The first term is simply the mass defined at infinity, and the second term is minus the Newtonian potential energy associated with building a spherical shell of mass and radius from individual particles, bringing them together from infinity. However, taking into account that on the Schwarzschild horizon while at the spatial infinity it is just , the Brown–York energy is monotonically decreasing with . Also, the first law of black hole mechanics for spherically symmetric black holes can be recovered by identifying with the internal energy [93, 94] . The thermodynamics of the Schwarzschild–anti-de-Sitter black holes was investigated in terms of the quasi-local quantities in [89] . Still considering to be the internal energy, the temperature, surface pressure, heat capacity, etc. are calculated (see subsection 13.3.1 ). The energy has also been calculated for the Einstein–Rosen cylindrical waves [92] . The energy is explicitly calculated for three different kinds of 2-spheres in the slices (in the Boyer–Lindquist coordinates) of the slow rotation limit of the Kerr black hole spacetime with the flat space reference [254] . These surfaces are the surfaces (such as the outer horizon), spheres whose intrinsic metric (in the given slow rotation approximation) is of a metric sphere of radius with surface area , and the ergosurface (i. e. the outer boundary of the ergosphere). The slow rotation approximation is defined such that , where is the typical spatial measure of the 2-surface. In the first two cases the angular momentum parameter enters the energy expression only in the order. In particular, the energy for the outer horizon is , which is twice the irreducible mass of the black hole. An interesting feature of this calculation is that the energy cannot be calculated for the horizon directly, because, as we noted in the previous point, the horizon itself cannot be isometrically embedded into a flat 3-space if the angular momentum parameter exceeds the irreducible mass [327] . The energy for the ergosurface is positive, as for the other two kinds of surfaces. The spacelike infinity limit of the charges interpreted as the energy, spatial momentum and spatial angular momentum are calculated in [92] (see also [171] ). Here the flat space reference configuration and the asymptotic Killing vectors of the spacetime are used, and the limits coincide with the standard ADM energy, momentum and spatial angular momentum. The analogous calculation for the centre-of-mass is given in [41] . It is shown that the corresponding large sphere limit is just the centre-of-mass expression of Beig and O Murchadha [46] . Here the centre-of-mass integral in terms of a charge integral of the curvature is also given. Although the prescription for the reference configuration by Hawking and Horowitz cannot be imposed for a general timelike 3-boundary (see subsection 10.1.8 ), asymptotically, when is pushed out to infinity, this prescription can be used, and coincides with the prescription of Brown and York. Choosing the background metric to be the anti-de-Sitter one, Hawking and Horowitz [171] calculated the limit of the quasi-local energy, and they found it to tend to the Abbott–Deser energy. (For the spherically symmetric, Schwarzschild–anti-de-Sitter case see also [89] .) In [90] the null infinity limit of the integral of was calculated both for the lapses generating asymptotic time translations and supertranslations at the null infinity, and the fleet of observers was chosen to tend to the BMS translation. In the former case the Bondi–Sachs energy, in the latter case Geroch’s supermomenta are recovered. These calculations are based directly on the Bondi form of the spacetime metric, and do not use the asymptotic solution of the field equations. In a slightly different formulation Booth and Creighton calculated the energy flux of outgoing gravitational radiation [74] (see also subsection 13.1 ) and they recovered the Bondi–Sachs mass-loss. However, the calculation of the small sphere limit based on the flat space reference configuration gave strange results [242] . While in non-vacuum the quasi-local energy is the expected , in vacuum it is proportional to instead of the Bel–Robinson ‘energy’ . (Here and are, respectively, the conformal electric and conformal magnetic curvatures, and plays a double role: It defines the 2-sphere of radius [as is usual in the small sphere calculations], and defines the fleet of observers on the 2-sphere.) On the other hand, the special light cone reference used in [91, 242] reproduces the expected result in non-vacuum, and yields in vacuum. The light cone reference was shown to work in the large sphere limit near the null and spatial infinities of asymptotically flat, and near the infinity of asymptotically anti-de-Sitter spacetimes [241] . Namely, the Brown–York quasi-local energy expression with this null cone reference term tends to the Bondi–Sachs, the ADM and Abbott–Deser energies, respectively. The supermomenta of Geroch at null infinity can also be recovered in this way. The proof is simply a demonstration of the fact that this light cone and the flat space prescriptions for the subtraction term have the same asymptotic structure up to order . This choice seems to work properly only in the asymptotics, because for small ellipsoids in the Minkowski spacetime this definition yields non-zero energy and for small spheres in vacuum it does not yield the Bel–Robinson ‘energy’ [243] .10.1.8 Other prescriptions for the reference configuration
As we noted above, Hawking, Horowitz and Hunter [171, 172] defined their reference configuration by embedding the Lorentzian 3-manifold isometrically into some given Lorentzian spacetime, e. g. into the Minkowski spacetime. (See also [154] .) However, for the given intrinsic 3-metric and the embedding 4–geometry the corresponding Gauss and Codazzi–Mainardi equations form a system of equations for the six components of the extrinsic curvature [93] . Thus, in general, this is a highly overdetermined system, and hence it may be expected to have a solution only in exceptional cases. However, even if such an embedding existed, then even the small perturbations of the intrinsic metric would break the conditions of embeddability. Therefore, in general this prescription for the reference configuration can work only if the 3-surface is ‘pushed out to infinity’ but does not work for finite 3-surfaces [93] . To rule out the possibility that the Brown–York energy can be non-zero even in Minkowski spacetime (on 2-surfaces in the boosted flat data set), Booth and Mann [75] suggested to embed isometrically into a reference spacetime (mostly into the Minkowski spacetime) instead of a spacelike slice of it, and to map the evolution vector field of the dynamics, tangent to , to a vector field in such that and . Here is a diffeomorphism mapping an open neighbourhood of in into such that , the restriction of to , is an isometry, and denotes the Lie derivative of along . This condition might be interpreted as some local version of that of Hawking, Horowitz and Hunter. However, Booth and Mann did not investigate the existence or the uniqueness of this choice.10.2 Kijowski’s approach
10.2.1 The role of the boundary conditions
In the Brown–York approach the leading principle was the claim to have a well defined variational principle. This led them to modify the Hilbert action to the trace- -action and the boundary condition that the induced 3-metric on the boundary of the domain of the action is fixed. However, as stressed by Kijowski [231, 233] , the boundary conditions have much deeper content. For example in thermodynamics the different definitions of the energy (internal energy, enthalpy, free energy, etc.) are connected with different boundary conditions. Fixing the pressure corresponds to enthalpy, but fixing the temperature to free energy. Thus the different boundary conditions correspond to different physical situations, and, mathematically, to different phase spaces. Therefore, to relax the a priori boundary conditions Kijowski abandoned the variational principle and concentrated on the equations of motions. However, to treat all possible boundary conditions on an equal footing he used the enlarged phase space of Tulczyjew (see for example [233] ). The boundary condition of Brown and York is only one of the possible boundary conditions.According to this view the quasi-local energy is similar to of ( 6 ), rather than to the charges which are connected somehow to some ‘absolute’ element of the spacetime structure.
This phase space is essentially , the cotangent bundle of the tangent bundle of the configuration manifold , endowed with the natural symplectic structure, and can be interpreted as the collection of quadruples . The usual Lagrangian (or velocity) phase space and the Hamiltonian (or momentum) phase space are special submanifolds of .
10.2.2 The analysis of the Hilbert action and the quasi-local internal and free energies
Starting with the variation of Hilbert’s Lagrangian (in fact, the corresponding Hamilton–Jacobi principal function on a domain above), and defining the Hamiltonian by the standard Legendre transformation on the typical compact spacelike 3-manifold and its boundary too, Kijowski arrived at a variation formula involving the value on of the variation of the canonical momentum, , conjugate to . (Apart from a numerical coefficient and the subtraction term, this is essentially the surface stress-energy tensor given by ( 64 ).) Since, however, it is not clear whether or not the initial+boundary value problem for the Einstein equations with fixed canonical momenta (i. e. extrinsic curvature) is well posed, he did not consider the resulting Hamiltonian as the appropriate one, and made further Legendre transformations on the boundary . The first Legendre transformation that he considered gave a Hamiltonian whose variation involves the variation of the induced 2-metric on and the parts and of the canonical momentum above. Explicitly, with the notations of the previous subsection, the latter two are and , respectively. ( is the de-densitized .) Then, however, the lapse and the shift on the boundary will not be independent: As Kijowski shows they are determined by the boundary conditions for the 2-metric and the freely specifiable parts and of the canonical momentum . Then, to define the ‘quasi-symmetries’ of the 2-surface, Kijowski suggests to embed first the 2-surface isometrically into an hyperplane of the Minkowski spacetime, and then define a world tube by dragging this 2-surface along the integral curves of the Killing vectors of the Minkowski spacetime. For example, to define the ‘quasi time translation’ of the 2-surface in the physical spacetime we must consider the time translation in the Minkowski spacetime of the 2-surface embedded in the hyperplane. This world tube gives an extrinsic curvature and vector potential . Finally, Kijowski’s choice for and is just and , respectively. In particular, to define the ‘quasi time translation’ he takes and , because this choice yields zero shift and constant lapse with value 1. The corresponding quasi-local quantity, the Kijowski energy, is(75) |
(76) |
10.3 The expression of Epp
10.3.1 The general form of Epp’s expression
The Brown–York energy expression, based on the original flat space reference, has the highly undesirable property that it gives non-zero energy even in the Minkowski spacetime if the fleet of observers on the spherical is chosen to be radially accelerating (see the second paragraph in subsection 10.1.7 ). Thus it would be a legitimate aim to reduce this extreme dependence of the quasi-local energy on the choice of the observers. One way of doing this is to formulate the quasi-local quantities in terms of boost-gauge invariant objects. Such a boost-gauge invariant geometric object is the length of the mean extrinsic curvature vector of subsection 4.1.2 , which, in the notations of the present section, is . If is spacelike or null, then this square root is real, and (apart from the reference term in ( 70 )) in the special case it reduces to -times the surface energy density of Brown and York. This observation lead Epp to suggest(77) |
10.3.2 The definition of the reference configuration
The subtraction term in ( 77 ) is defined through an isometric embedding of into some reference spacetime instead of a 3-space. This spacetime is usually Minkowski or anti-de-Sitter spacetime. Since the 2-surface data consist of the metric, the two extrinsic curvatures and the -gauge potential, for given and ambient spacetime the conditions of the isometric embedding form a system of six equations for eight quantities, namely for the two extrinsic curvatures and the gauge potential (see subsection 4.1.2 , and especially equation ( 20 )-( 21 )). Therefore, even a naive function counting argument suggests that the embedding exists, but is not unique. To have uniqueness additional conditions must be imposed. However, since is a gauge field, one condition might be a gauge fixing in the normal bundle, and Epp’s suggestion is to require that the curvature of the connection 1-form in the reference spacetime and in the physical spacetime be the same [129] . Or, in other words, not only the intrinsic metric of is required to be preserved in the embedding, but the whole curvature of the connection as well. In fact, in the connection on the spinor bundle both the Levi-Civita and the connection coefficients appear on an equal footing. (Recall that we interpreted the connection to be a part of the universal structure of .) With this choice of the reference configuration depends not only on the intrinsic 2-metric of , but on the connection on the normal bundle as well. Suppose that is a 2-surface in such that with , and, in addition, can be embedded into the flat 3-space with . Then there is a boost gauge (the ‘quasi-local rest frame’) in which coincides with the Brown–York energy in the particular boost-gauge for which . Consequently, every statement stated for the latter is valid for , and every example calculated for is an example for as well [129] . A clear and careful discussion of the potential alternative choices for the reference term, especially their potential connection with the angular momentum, is also given there.10.3.3 The various limits
First, it should be noted that Epp’s quasi-local energy is vanishing in Minkowski spacetime for any 2-surface, independently of any fleet of observers. In fact, if is a 2-surface in Minkowski spacetime, then the same physical Minkowski spacetime defines the reference spacetime as well, and hence . For round spheres in the Schwarzschild spacetime it yields the result that gave. In particular, for the horizon it is (instead of ), and at infinity it is [129] . Thus, in particular, is also monotonically decreasing with in Schwarzschild spacetime. Epp calculated the various limits of his expression too [129] . In the large sphere limit near spatial infinity he recovered the Ashtekar–Hansen form of the ADM energy, at future null infinity the Bondi–Sachs energy. The technique that is used in the latter calculations is similar to that of [90] . In non-vacuum in the small sphere limit reproduces the standard result, but the calculations for the vacuum case are not completed. The leading term is still probably of order , but its coefficient has not been calculated. Although in these calculations plays the role only of fixing the 2-surfaces, as a result we got energy seen by the observer instead of mass. It is this reason why is considered to be energy rather than mass. In the asymptotically anti-de-Sitter spacetime (with the anti-de-Sitter spacetime as the reference spacetime) gives zero. This motivated Epp to modify his expression to recover the mass parameter of the Schwarzschild–anti-de-Sitter spacetime at the infinity. The modified expression is, however, not boost-gauge invariant. Here the potential connection with the AdS/CFT correspondence is also discussed (see also [32] ).10.4 The expression of Liu and Yau
10.4.1 The Liu–Yau definition
Let be a spacelike topological 2-sphere in spacetime such that the metric has positive scalar curvature. Then by the embedding theorem there is a unique isometric embedding of into the flat 3-space, and this embedding is unique. Let be the trace of the extrinsic curvature of in this embedding, which is completely determined by and is necessarily positive. Let and be the trace of the extrinsic curvatures of in the physical spacetime corresponding to the outward pointing unit spacelike and future pointing timelike normals, respectively. Then Liu and Yau define their quasi-local energy in [245] by(78) |
(79) |
10.4.2 The main properties of
The most important property of the quasi-local energy definition ( 78 ) is its positivity. Namely [245] , let be a compact spacelike hypersurface with smooth boundary , consisting of finitely many connected components , . . . , such that each of them has positive intrinsic curvature. Suppose that the matter fields satisfy the dominant energy condition on . Then is strictly positive unless the domain of dependence is flat. In this case is connected. The proof is based on the use of Jang’s equation [211] , by means of which the general case can be reduced to the results of Shi and Tam in the time symmetric case [326] , stated in subsection 10.1.7 . (See also [380] .) If is an apparent horizon, i. e. , then is just the integral of . Then by the Minkowski inequality for the convex surfaces in the flat 3-space (see for example [363] ) one has i. e. it is not less than twice the irreducible mass of the horizon. For round spheres coincides with , and hence it does not reduce to the standard round sphere expression ( 27 ). In particular, for the event horizon of the Schwarzschild black hole it is . Although the strict mathematical analysis is still lacking, probably reproduces the correct large sphere limits in asymptotically flat spacetime (ADM and Bondi–Sachs energies), because the difference between the Brown–York, Epp and Kijowski–Liu–Yau definitions disappear asymptotically. However, can be positive even if is in the Minkowski spacetime. In fact, for given intrinsic metric on (with positive scalar curvature) can be embedded into the flat ; this embedding is unique, and the trace of the extrinsic curvature is determined by . On the other hand, the isometric embedding of in the Minkowski spacetime is not unique: The equations of the embedding (i. e. the Gauss, the Codazzi–Mainardi and the Ricci equations) form a system of six equations for the six components of the two extrinsic curvatures and and the two components of the gauge potential . Thus, even if we impose a gauge condition for the connection 1-form , we have only six equations for the seven unknown quantities, leaving enough freedom to deform in the Minkowski spacetime to get positive Kijowski–Liu–Yau energy. Indeed, specific 2-surfaces in the Minkowski spacetime are given in [?] for which .11 Towards a Full Hamiltonian Approach
The Hamilton–Jacobi method is only one possible strategy to define the quasi-local quantities in a large class of approaches, called the Hamiltonian or canonical approaches. Thus there is a considerable overlap between the various canonical methods, and hence the cutting of the material into two parts (section 10 and section 11 ) is, in some sense, artifical. In the previous section we reviewed those approaches that are based on the analysis of the action, while in the present we discuss those that are based primarily of the analysis of the Hamiltonian in the spirit of Regge and Teitelboim [304] . By a full Hamiltonian analysis we mean a detailed study of the structure of the quasi-local phase space, including the constraints, the smearing fields, the symplectic structure and the Hamiltonian itself, according to the standard or some generalized Hamiltonian scenarios, in the traditional 3+1 or in the fully Lorentz-covariant form, or even in the 2+2 form, using the metric or triad/tetrad variables (or even the Weyl or Dirac spinors). In the literature of canonical general relativity (at least in the asymptotically flat context) there are examples for all these possibilities, and we report on the quasi-local investigations on the basis of the decomposition they use. Since the 2+2 decomposition of the spacetime is less known, we also summarize its basic idea.In fact, Kijowski’s results could have been presented here, but the technique that he uses may justify their inclusion in the previous section.
11.1 The 3+1 approaches
There is a lot of literature on the canonical formulation of general relativity both in the traditional ADM and the Møller tetrad (or, recently, the closely related complex Ashtekar) variables. Thus it is quite surprising how little effort has been spent to systematically quasi-localize them. One motivation for the quasi-localization of the ADM–Regge–Teitelboim analysis came from the need for the microscopic understanding of black hole entropy [31, 30, 96] : What are the microscopic degrees of freedom behind the phenomenological notion of black hole entropy? Since the aim of the present paper is to review the construction of the quasi-local quantities in classical general relativity, we discuss only the classical 2-surface observables by means of which the ‘quantum edge states’ on the black hole event horizons were intended to be constructed.11.1.1 The 2-surface observables
If , the 3-manifold on which the ADM canonical variables , are defined, has a boundary , then the usual vacuum constraints(80) |
(81) |
(82) |
(83) |
Here we concentrate only on the genuine, finite boundary of . The analysis is straightforward even in the presence of ‘boundaries at infinity’ at the asymptotic ‘ends’ of asymptotically flat .
11.2 Approaches based on the double-null foliations
11.2.1 The 2+2 decomposition
The decomposition of the spacetime in a 2+2 way with respect to two families of null hypersurfaces is as old as the study of gravitational radiation and the concept of the characteristic initial value problem (see for example [311, 292] ). The basic idea is that we foliate an open subset of the spacetime by a 2-parameter family of (e. g. closed) spacelike 2-surfaces: If is the typical 2-surface, then this foliation is defined by a smooth embedding . Then, keeping fixed and varying , or keeping fixed and varying , respectively, defines two 1-parameter families of hypersurfaces and . Requiring one (or both) of the hypersurfaces to be null, we get a so-called null (or double-null, respectively) foliation of . (In subsection 4.1.8 we required the hypersurfaces to be null only for the special value of the parameters.) As is well known, because of the conjugate points, in the null or double null cases the foliation can be well defined only locally. For fixed and the prescription defines a curve through in , and hence a vector field tangent everywhere to on . The Lie bracket of and the analogously defined is zero. There are several inequivalent ways of introducing coordinates or rigid frame fields on , which are fit naturally to the null or double null foliation , in which the (vacuum) Einstein equations and Bianchi identities take a relatively simple form [311, 148, 119, 332, 364, 175, 161, 80, 184] . Defining the ‘time derivative’ to be the Lie derivative, for example, along the vector field , the Hilbert action can be rewritten according to the 2+2 decomposition. Then the 2+2 form of the Einstein equations can be derived from the corresponding action as the Euler–Lagrange equations provided the fact that the foliation is null is imposed only after the variation has made. (Otherwise, the variation of the action with respect to the less than ten nontrivial components of the metric would not yield all the 10 Einstein equations.) One can form the corresponding Hamiltonian, in which the null character of the foliation should appear as a constraint. Then the formal Hamilton equations are just the Einstein equations in their 2+2 form [119, 364, 175, 184] . However, neither the boundary terms in this Hamiltonian nor the boundary conditions that could ensure its functional differentiability were considered. Therefore, this Hamiltonian can be ‘correct’ only up to boundary terms. Such a Hamiltonian was used by Hayward [175, 178] as the basis of his quasi-local energy expression discussed already in subsection 6.3 . (A similar energy expression was derived by Ikumi and Shiromizi [205] , starting with the idea of the ‘freely falling 2-surfaces’.)11.2.2 The 2+2 quasi-localization of the Bondi–Sachs mass-loss
As we mentioned in subsection 6.1.3 , this double-null foliation was used by Hayward [177] to quasi-localize the Bondi–Sachs mass-loss (and mass-gain) by using the Hawking energy. Thus we do not repeat the review of his results here. Yoon investigated the vacuum field equations in a coordinate system based on a null 2+2 foliation. Thus one family of hypersurfaces was (outgoing) null, e. g. , but the other was timelike, say . The former defined a foliation of the latter in terms of the spacelike 2-surfaces . Yoon found [381, 382] a certain 2-surface integral on , denoted by , for which the difference , , could be expressed as a flux integral on the portion of the timelike hypersurface between and . In general this flux does not have a definite sign, but Yoon showed that asymptotically, when is ‘pushed out to null infinity’ (i. e. in the limit in an asymptotically flat spacetime), it becomes negative definite. In fact, ‘renormalizing’ by a subtraction term, tends to the Bondi energy, and the flux integral tends to the Bondi mass-loss between the cuts and [381, 382] . These investigations were extended for other integrals in [383, 384] , which are analogous to spatial momentum and angular momentum. However, all these integrals, including above, depend not only on the geometry of the spacelike 2-surface but on the 2+2 foliation on an open neighbourhood of too.11.3 The covariant approach
11.3.1 The covariant phase space methods
The traditional ADM approach to conserved quantities and the Hamiltonian analysis of general relativity is based on the 3+1 decomposition of fields and geometry. Although the results and the content of a theory may be covariant even if their form is not, the manifest spacetime covariance of a formalism may help to find the (spacetime covariant) observables and conserved quantities, boundary conditions, etc. easily. No a posteriori spacetime interpretation of the results is needed. Such a spacetime–covariant Hamiltonian formalism was initiated by Nester [269, 272] . His basic idea is to use (tensor or Dirac spinor valued) differential forms as the basic field variables on the spacetime manifold . Thus his phase space is the collection of fields on the 4-manifold , endowed with the (generalized) symplectic structure of Kijowski and Tulczyjew [233] . He derives the field equations from the Lagrangian 4-form, and for a fixed spacetime vector field finds a Hamiltonian 3-form whose integral on a spacelike hypersurface takes the form(84) |
11.3.2 The covariant quasi-local Hamiltonians
The quasi-local Hamiltonian for a large class of geometric theories, allowing torsion and non-metricity of the connection, was investigated by Chen, Nester and Tung [106, 104, 274] in the covariant approach of Nester above [269, 272] . Starting with a Lagrangian 4-form for a first order formulation of the theory and an arbitrary vector field , they determine the general form of the Hamiltonian 3-form , including the boundary 2-form . However, in the variation of the corresponding Hamiltonian there will be boundary terms in general. To cancel them, the boundary 2-form has to be modified. Introducing an explicit reference field and canonical momentum (which are solutions of the field equations), Chen, Nester and Tung suggest (in the differential form notation) either of the two 4-covariant boundary 2-forms(85) |
(86) |
11.3.3 Pseudotensors and quasi-local quantities
As we discussed briefly in section 3.3.1 , many, apparently different pseudotensors and -gauge dependent energy-momentum density expressions can be recovered from a single differential form defined on the bundle of linear frames over the spacetime manifold: The corresponding superpotentials are the pull backs to of the various forms of the Nester–Witten 2-from from along the various local sections of the bundle [138, 256, 336, 337] . Thus the different pseudotensors are simply the gauge dependent manifestations of the same geometric object on the bundle in the different gauges. Since, however, is the unique extension of the Nester–Witten 2-form on the principal bundle of normalized spin frames (given in equation ( 12 )), and the latter has been proven to be connected naturally to the gravitational energy-momentum, the pseudotensors appear to describe the same physics as the spinorial expressions, though in a slightly old fashioned form. That this is indeed the case was demonstrated clearly by Chang, Nester and Chen [101, 105, 274] , by showing an intimate connection between the covariant quasi-local Hamiltonian expressions and the pseudotensors. Writing the Hamiltonian in the form of the sum of the constraints and a boundary term, in a given coordinate system the integrand of this boundary term may be the superpotential of any of the pseudotensors. Then the requirement of the functional differentiability of gives the boundary conditions for the basic variables at . For example, for the Freud superpotential (for Einstein’s pseudotensor) what is fixed on the boundary is a certain piece of .12 Constructions for Special Spacetimes
12.1 The Komar integral for spacetimes with Killing vectors
Although the Komar integral (and, in general, the linkage ( 16 ) for some ) does not satisfy our general requirements discussed in subsection 4.3.1 , and it does not always give the standard values in specific situations (see for example the ‘factor-of-two anomaly’ or the examples below), in the presence of a Killing vector the Komar integral, built from the Killing field, could be a very useful tool in practice. (For Killing fields the linkage reduces to the Komar integral for any .) One of its most important properties is that in vacuum depends only on the homology class of the 2-surface (see for example [369] ): If and are any two 2-surfaces such that for some compact three dimensional hypersurface on which the energy-momentum tensor of the matter fields is vanishing, then . In particular, the Komar integral for the static Killing field in the Schwarzschild spacetime is the mass parameter of the solution for any 2-surface surrounding the black hole, but it is zero if does not. On the other hand [354] , the analogous integral in the Reissner–Nordstrom spacetime on a metric 2-sphere of radius is , which deviates from the generally accepted round-sphere value . Similarly, in Einstein’s static universe for the spheres of radius in a hypersurface is zero instead of the round sphere result , where is the energy density of the matter and is the cosmological constant.12.2 The effective mass of Kulkarni, Chellathurai and Dadhich for the Kerr spacetime
The Kulkarni–Chellathurai–Dadhich [237] effective mass for the Kerr spacetime is obtained from the Komar integral (i. e. the linkage with ) using a hypersurface orthogonal vector field instead of the Killing vector of stationarity. The vector field is defined to be , where is the Killing vector of axis-symmetry and the function is . This is timelike outside the horizon, it is the asymptotic time translation at infinity and coincides with the null tangent on the event horizon. On the event horizon it yields , while in the limit it is the mass parameter of the solution. The effective mass is computed for the Kerr–Newman spacetime in [103] .12.3 The Katz–Lynden-Bell–Israel energy for static spacetimes
Let be a hypersurface-orthogonal timelike Killing vector field, a spacelike hypersurface to which is orthogonal, and . Let be the set of those points of where the length of the Killing field is the value , i. e. are the equipotential surfaces in , and let be the set of those points where the magnitude of is not greater than . Suppose that is compact and connected. Katz, Lynden-Bell and Israel [226] associate a quasi-local energy to the 2-surfaces as follows. Suppose that the matter fields can be removed from and can be concentrated into a thin shell on in such a way that the space inside be flat but the geometry outside remain the same. Then, denoting the (necessarily distributional) energy-momentum tensor of the shell by and assuming that it satisfies the weak energy condition, the total energy of the shell, , is positive. Here is the future directed unit normal to . Then, using the Einstein equations, the energy of the shell can be rewritten in terms of geometric objects on the 2-surface as(87) |
13 Applications in General Relativity
In this part we give a very short review of some of the potential applications of the paradigm of quasi-locality in general relativity. This part of the review is far from being complete, and our claim here is not to discuss the problems considered in detail, but rather to give a collection of problems that are (effectively or potentially) related to quasi-local ideas, tools, notions etc. In some of these problems the various quasi-local expressions and techniques have been used successfully, but others may provide new and promising areas of their application.13.1 Calculation of tidal heating
According to astronomical observations, there is an intensive volcanism on the moon Io of Jupiter. One possible explanation of this phenomenon is that Jupiter is heating Io via the gravitational tidal forces (like the Moon, whose gravitational tidal forces raise the ocean’s tides on the Earth). To check whether this could be really the case, one must be able to calculate how much energy is pumped into Io. However, gravitational energy (both in Newtonian theory and in general relativity) is only ambiguously defined (and hence cannot be localized), while the phenomena mentioned above cannot depend on the mathematics that we use to describe them. The first investigations intended to calculate the tidal work (or heating) of a compact massive body were based on the use of the various gravitational pseudotensors [303, 132] . It has been shown that although in the given (slow motion and isolated body) approximation the interaction energy between the body and its companion is ambiguous, the tidal work that the companion does on the body via the tidal forces is not. This is independent both of the gauge conditions [303] and the actual pseudotensor (Einstein, Møller, Bergmann or Landau–Lifshitz) [132] . Recently, these calculations were repeated using quasi-local concepts by Booth and Creighton [74] . They calculated the time derivative of the Brown–York energy, given by ( 69 )-( 70 ). Assuming the form of the metric used in the pseudotensorial calculations, for the tidal work they recovered the gauge invariant expressions obtained in [303, 132] . In these approximate calculations the precise form of the boundary conditions (or reference configurations) is not essential, because the results obtained by using different boundary conditions deviate from each other only in higher order.13.2 Geometric inequalities for black holes
13.2.1 On the Penrose inequality
To rule out a certain class of potential counterexamples to the (weak) cosmic censorship hypothesis [289] , Penrose derived an inequality that any asymptotically flat initial data set with (outermost) apparent horizon must satisfy [291] : The ADM mass of the data set cannot be less than the so-called irreducible mass of the horizon, (see also [152, 363] ). This inequality has been proven for spherically symmetric spacetimes (using the Hawking energy) [252] (see also [388] ), for static black holes (using the Penrose mass, as we mentioned in subsection 7.2.5 ) [357, 358] , and for the perturbed Reissner–Nordstrom spacetimes [219] (see also [220] ). Although the original specific potential counterexample has been shown not to violate the Penrose inequality [153] , the inequality has not been proven for general data set. (For the limitations of the proof of the Penrose inequality for the area of a trapped surface and the Bondi mass at past null infinity [249] see [62] .) If the inequality were true, then this would be a strengthened version of the positive mass theorem, providing a positive lower bound for the ADM mass. On the other hand, for time symmetric data sets the Penrose inequality has been proven, even in the presence of more than one black hole. The proof is based on the use of some quasi-local energy expression, mostly of Geroch or of Hawking. First it is shown that these expressions are monotonic along the vector field of a special foliation of the time symmetric initial hypersurface (see subsections 6.1.3 and 6.2 , and also [139] ), and then the global existence of such a foliation between the apparent horizon and the 2-sphere at infinity is proven. The first complete proof of the latter was given by Huisken and Ilmanen [201, 202] . (Recently Bray used a conformal technique to give an alternative proof [84, 85, 86] .) A more general form of the conjecture, containing electric charge of the black hole, was formulated by Gibbons [152] : The ADM mass is claimed not to be exceeded by . Although the weaker form of the inequality, the so-called Bogomolny inequality , has been proven (under assumptions on the matter content see for example [156, 352, 248, 155, 263, 152] ), Gibbons’ inequality for the electric charge has been proven for special cases (for spherically symmetric spacetimes see for example [181] ), and for time symmetric initial data sets using Geroch’s inverse mean curvature flow [212] . As a consequence of the results of [201, 202] the latter has become a complete proof. However, this inequality does not seem to work in the presence of more than one black hole: For a time symmetric data set describing nearly extremal Reissner–Nodstrom black holes can be greater than the ADM mass, where is either the area of the outermost marginally trapped surface [?] , or the sum of the areas of the individual black hole horizons [120] . On the other hand, the weaker inequality derived from the cosmic censorship assumption, does not seem to be violated even in the presence of more than one black hole. If in the final state of gravitational collapse the black hole has not only electric charge but angular momentum as well, then the geometry is described by the Kerr–Newman solution. Expressing the mass parameter of the solution (which is just -times the ADM mass) in terms of the irreducible mass of the horizon, the electric charge and the angular momentum , we arrive at the more general form(88) |
I am grateful to Sergio Dain for pointing out this to me.
13.2.2 On the hoop conjecture
In connection with the formation of black holes and the weak cosmic censorship hypothesis another geometric inequality has also been formulated. This is the hoop conjecture of Thorne [350, 259] , saying that ‘black holes with horizons form when and only when a mass gets compacted into a region whose circumference in every direction is ’ (see also [135, 373] ). Mathematically, this conjecture is not precisely formulated: Neither the mass nor the notion of the circumference is well defined. In certain situations the mass might be the ADM or the Bondi mass, but might be the integral of some locally defined ‘mass density’ as well [135, 34] . The most natural formulation of the hoop conjecture would be based on some spacelike 2-surface and some reasonable notion of the quasi-local mass, and the trapped nature of the surface would be characterized by the mass and the ‘circumference’ of . In fact, for round spheres outside the outermost trapped surface and the standard round sphere definition of the quasi-local energy ( 26 ) one has , where we used the fact that is an areal radius (see subsection 4.2.1 ). If, however, is not axis-symmetric then there is no natural definition (or, there are several inequivalent ‘natural’ definitions) of the circumference of . For the investigations of the hoop conjecture in the Gibbons–Penrose spacetime of the collapsing thin matter shell see [35, 34, 362, 284] , and for colliding black holes see [387] .13.3 Quasi-local laws of black hole dynamics
13.3.1 Quasi-local thermodynamics of black holes
Black holes are usually introduced in asymptotically flat spacetimes [167, 168, 170, 369] , and hence it was natural to derive the formal laws of black hole mechanics/thermodynamics in the asymptotically flat context (see for example [33, 49, 50] , and for a recent review see [374] ). The discovery of the Hawking radiation [169] showed that the laws of black hole thermodynamics are not only analogous to the laws of thermodynamics, but black holes are genuine thermodynamical objects: The black hole temperature is a physical temperature, that is -times the surface gravity, and the entropy is a physical entropy, -times the area of the horizon (in the traditional units with the Boltzmann constant , speed of light , Newton’s gravitational constant and Planck’s constant ) (see also [372] ). Apparently, the detailed microscopic (quantum) theory of gravity is not needed to derive the black hole entropy, and it can be derived even from the general principles of a conformal field theory on the horizon of the black holes [97, 98, 99, 282, 100] . However, black holes are localized objects, thus one must be able to describe their properties and dynamics even at the quasi-local level. Nevertheless, beyond this rather theoretic claim, there are pragmatic reasons that force us to quasi-localize the laws of black hole dynamics. In particular, it is well known that the Schwarzschild black hole, fixing its temperature at infinity, has negative heat capacity. Similarly, in an asymptotically anti-de-Sitter spacetime fixing the black hole temperature via the normalization of the timelike Killing vector at infinity is not justified because there is no such physically distinguished Killing field (see [89] ). These difficulties lead to the need of a quasi-local formulation of black hole thermodynamics. In [89] Brown, Creighton and Mann investigated the thermal properties of the Schwarzschild–anti-de-Sitter black hole. They used the quasi-local approach of Brown and York to define the energy of the black hole on a spherical 2-surface outside the horizon. Identifying the Brown–York energy with the internal (thermodynamical) energy and (in the units) -times the area of the event horizon with the entropy, they calculated the temperature, surface pressure and heat capacity. They found that these quantities do depend on the location of the surface . In particular, there is a critical value such that for temperatures greater than there are two black hole solutions, one with positive and one with negative heat capacity, but there are no Schwarzschild–anti-de-Sitter black holes with temperature less than . In [117] the Brown–York analysis is extended to include dilaton and Yang–Mills fields, and the results are applied to stationary black holes to derive the first law of black hole thermodynamics. The so-called Noether charge formalism of Wald [371] and Iyer and Wald [209] can be interpreted as a generalization of the Brown–York approach from general relativity to any diffeomorphism invariant theory to derive quasi-local quantities [210] . However, this formalism gave a general expression for the black hole entropy as well: That is the Noether charge derived from the Hilbert Lagrangian corresponding to the null normal of the horizon, and explicitly this is still -times the area of the horizon. (For some recent related works see for example [145, 183] ). There is an extensive literature of the quasi-local formulation of the black hole dynamics and relativistic thermodynamics in the spherically symmetric context (see for example [180, 182, 181, 185] and for non-spherically symmetric cases [264, 184, 72] ). However, one should see clearly that while the laws of black hole thermodynamics above refer to the event horizon, which is a global concept in the spacetime, the subject of the recent quasi-local formulations is to describe the properties and the evolution of the so-called trapping horizon, which is a quasi-locally defined notion. (On the other hand, the investigations of [178, 176, 179] are based on energy and angular momentum definitions that are gauge dependent. See also subsections 4.1.8 and 6.3 .)13.3.2 On the isolated horizons
The idea of the isolated horizons (more precisely, the gradually more restrictive notions of the non-expanding, the weakly isolated and isolated horizons, and the special weakly isolated horizon called the rigidly rotating horizons) is to generalize the notion of Killing horizons by keeping their basic properties without the existence of any Killing vector in general. (For a recent review see [18] and references therein, especially [20, 17] .) The phase space for asymptotically flat spacetimes containing an isolated horizon is based on a 3-manifold with an asymptotic end (or finitely many such ends) and an inner boundary. The boundary conditions on the inner boundary are determined by the precise definition of the isolated horizon. Then, obviously, the Hamiltonian will be the sum of the constraints and boundary terms, corresponding both to the ends and the horizon. Thus, by the appearance of the boundary term on the inner boundary makes the Hamiltonian partly quasi-local. It is shown that the condition of the Hamiltonian evolution of the states on the inner boundary along the evolution vector field is precisely the first law of black hole mechanics [20, 17] . Booth [73] applied the general idea of Brown and York to a domain whose boundary consists not only of two spacelike submanifolds and and a timelike one , but a further, internal boundary as well, which is null. Thus he made the investigations of the isolated horizons fully quasi-local. Therefore, the topology of and is , and the inner (null) boundary is interpreted as (a part of) a non-expanding horizon. Then to have a well defined variational principle on , the Hilbert action had to be modified by appropriate boundary terms. However, requiring to be a so-called rigidly rotating horizon, the boundary term corresponding to and the allowed variations are considerably restricted. This made it possible to derive the ‘first law of rigidly rotating horizon mechanics’ quasi-locally, an analog of the first law of black hole mechanics. The first law for rigidly rotating horizons was also derived by Allemandi, Francaviglia and Raiteri in the Einstein–Maxwell theory [4] using their Regge–Teitelboim-like approach [137] . Another concept is the notion of dynamical horizon [24, 25] . This is a smooth spacelike hypersurface that can be foliated by a preferred family of marginally trapped 2-spheres. By an appropriate definition of the energy and angular momentum balance equations for these quantities, carried by gravitational waves, are derived. Isolated horizons are the asymptotic state of dynamical horizons.13.4 Entropy bounds
13.4.1 On Bekenstein’s bounds for the entropy
Having associated the entropy to the (spacelike cross section of the) event horizon, it is natural to expect the generalized second law (GSL) of thermodynamics to hold, i. e. the sum of the entropy of the matter and the black holes cannot decrease in any process. However, as Bekenstein pointed out, it is possible to construct thought experiments (e. g. the so-called Geroch process) in which the GSL is violated, unless a universal upper bound for the entropy-to-energy ratio for bounded systems exists [51, 52] . (For another resolution of the apparent contradiction to the GSL, based on the calculation of the buoyancy force in the thermal atmosphere of the black hole, see [367, 372] .) In traditional units this upper bound is given by , where and are, respectively, the total energy and entropy of the system, and is the radius of the sphere that encloses the system. It is remarkable that this inequality does not contain Newton’s constant, and hence it can be expected to be applicable even for non-gravitating systems. Although this bound is violated for several model systems, for a wide class of systems in Minkowski spacetime the bound does hold [280, 281, 279, 53] (see also [79] ). The Bekenstein bound has been extended for systems with electric charge by Zaslavskii [389] , and for rotating systems by Hod [195] (see also [54, 162] ). Although these bounds were derived for test bodies falling into black holes, interestingly enough these Bekenstein bounds hold for the black holes themselves provided the generalized Gibbons–Penrose inequality ( 88 ) holds: Identifying with and letting be a radius for which is not less than the area of the event horizon of the black hole, ( 88 ) can be rewritten in the traditional units as(89) |
13.4.2 On the holographic hypothesis
In the literature there is another kind of upper bound for the entropy for a localized system, the so-called holographic bound. The holographic principle [349, 334, 79] says that, at the fundamental (quantum) level, the state of any physical system located in a compact spatial domain is fully characterized by the degrees of freedom on the surface of the domain, analogously to the holography by means of which a three dimensional image is encoded into a two dimensional surface. Consequently, the number of physical degrees of freedom in the domain is connected with the area of the boundary of the domain instead of its volume: The number of physical degrees of freedom is just one-fourth of the area of the surface measured in Planck-area units . This expectation is formulated in the (spacelike) holographic entropy bound [79] : Let be a compact spacelike hypersurface with boundary . Then the entropy of the system in should satisfy . Formally, this bound can be obtained from the Bekenstein bound with the assumption that , i. e. that is not less than the Schwarzschild radius of . Also, as with the Bekenstein bounds, this inequality can be violated in specific situations. (See also [374, 79] .) On the other hand, there is another formulation of the holographic entropy bound, due to Bousso [78, 79] . Bousso’s so-called covariant entropy bound is much more quasi-local than the previous formulations, and is based on spacelike 2-surfaces and the null hypersurfaces determined by the 2-surfaces in the spacetime. Its classical version has been proved by Flanagan, Marolf and Wald [136] : If is an everywhere non-contracting (or non-expanding) null hypersurface with spacelike cuts and , then, assuming that the local entropy density of the matter is bounded by its energy density, the entropy flux through between the cuts and is bounded: . For a detailed discussion see [374, 79] .13.5 Quasi-local radiative modes of GR
In subsection 8.2.3 we discussed the properties of the Dougan–Mason energy-momenta, and we saw that, under the conditions explained there, the energy-momentum is vanishing iff is flat, and it is null iff is a pp-wave geometry with pure radiative matter, and that these properties of the domain of dependence are completely encoded into the geometry of the 2-surface . However, there is an important difference between these two statements: While in the former case we know the metric of , that is flat, in the second we know only that the geometry admits a constant null vector field, but we do not know the line element itself. Thus the question arises as whether the metric of is also determined by the geometry of even in the zero quasi-local mass case. In [342] it was shown that under the condition above there is a complex valued function on , describing the deviation of the anti-holomorphic and the holomorphic spinor dyads from each other, which plays the role of a potential for the curvature on . Then, assuming that is future and past convex and the matter is an N-type zero-rest-mass field, and the value of the matter field on determine the curvature of . Since the field equations for the metric of reduce to Poisson-like equations with the curvature as the source, the metric of is also determined by and on . Therefore, the (purely radiative) pp-wave geometry and matter field on are completely encoded in the geometry of and complex functions defined on , respectively, in complete agreement with the holographic principle of the previous subsection. As we saw in subsection 2.2.5 , the radiative modes of the zero-rest-mass-fields in Minkowski spacetime, defined by their Fourier expansion, can be characterized quasi-locally on the globally hyperbolic subset of the spacetime by the value of the Fourier modes on the appropriately convex spacelike 2-surface . Thus the two transversal radiative modes of them are encoded in certain fields on . On the other hand, because of the non-linearity of the Einstein equations it is difficult to define the radiative modes of general relativity. It could be done when the field equations become linear, i. e. near the null infinity, in the linear approximation and for pp-waves. In the first case the gravitational radiation is characterized on a cut of the null infinity by the -derivative of the asymptotic shear of the outgoing null hypersurface for which ; i. e. by a complex function on . It is remarkable that it is precisely this complex function which yields the deviation of the holomorphic and anti-holomorphic spin frames at the null infinity (see for example [347] ). The linear approximation of Einstein’s theory is covered by the analysis of subsection 2.2.5 , thus those radiative modes can be characterized quasi-locally, while for the pp-waves the result of [342] , reported above, gives such a quasi-local characterization in terms of a complex function measuring the deviation of the holomorphic and anti-holomorphic spin frames. However, the deviation of the holomorphic and anti-holomorphic structures on can be defined even for generic 2-surfaces in generic spacetimes too, which might yield the possibility of introducing the radiative modes quasi-locally in general.14 Summary: Achievements, Difficulties and Open Issues
In the previous sections we tried to give an objective review of the present state of the art. This section is, however, less positivistic: We close the present review by a critical discussion, evaluating those strategies, approaches etc. that are explicitly given and (at least in principle) applicable in any generic spacetime.14.1 On the Bartnik mass and the Hawking energy
Although in the literature the notions mass and energy are used almost synonymously, in the present review we have made a distinction between them. By energy we meant the time component of the energy-momentum four-vector, i. e. a reference frame dependent quantity, while by mass the length of the energy-momentum, i. e. an invariant. In fact, these two have different properties: The quasi-local energy (both for the matter fields and for gravity according to the Dougan–Mason definition) is vanishing precisely for the ‘ground state’ of the theory (i. e. for vanishing energy-momentum tensor in the domain of dependence and flatness of , respectively, see subsections 2.2.5 and 8.2.3 ). In particular, for configurations describing pure radiation (purely radiative matter fields and pp-waves, respectively) the energy is positive. On the other hand, the vanishing of the quasi-local mass does not characterize the ‘ground state’, rather that is equivalent just to these purely radiative configurations. The Bartnik mass is a natural quasi-localization of the ADM mass, and its monotonicity and positivity makes it a potentially very useful tool in proving various statements on the spacetime, because it fully characterizes the non-triviality of the finite Cauchy data by a single scalar. However, our personal opinion is that, just by its strict positivity for non-flat 3-dimensional domains, it overestimates the ‘physical’ quasi-local mass. In fact, if is a finite data set for a pp-wave geometry (i. e. a compact subset of the data set for a pp-wave metric), then it probably has an asymptotically flat extension satisfying the dominant energy condition with bounded ADM energy and no apparent horizon between and infinity. Thus while the Dougan–Mason mass of is zero, the Bartnik mass is strictly positive unless is trivial. Thus, this example shows that it is the procedure of taking the asymptotically flat extension that gives strictly positive mass. Indeed, one possible proof of the rigidity part of the positive energy theorem [23] (see also [338] ) is to prove first that the vanishing of the ADM mass implies, through the Witten equation, that the spacetime admits a constant spinor field, i. e. it is a pp-wave spacetime, and then that the only asymptotically flat spacetime that admits a constant null vector field is the Minkowski spacetime. Therefore, it is just the global condition of the asymptotic flatness that rules out the possibility of non-trivial spacetimes with zero ADM mass. Hence it would be instructive to calculate the Bartnik mass for a compact part of a pp-wave data set. It might also be interesting to calculate its small surface limit to see its connection with the local fields (energy-momentum tensor and probably the Bel–Robinson tensor). The other very useful definition is the Hawking energy (and its slightly modified version, the Geroch energy). Its advantage is its simplicity, calculability and monotonicity for special families of 2-surfaces, and it has turned out to be a very effective tool in practice in proving for example the Penrose inequality. The small sphere limit calculation shows that it is energy rather than mass, so in principle one should be able to complete this to an energy-momentum 4-vector. One possibility is ( 36 )-( 37 ), but, as far as we are aware, its properties have not been investigated. Unfortunately, although the energy can be defined for 2-surfaces with nonzero genus, it is not clear how the 4-momentum could be extended for such surfaces. Although the Hawking energy is a well defined 2-surface observable, it has not been linked to any systematic (Lagrangian or Hamiltonian) scenario. Perhaps it does not have any such interpretation, and it is simply a natural (but in general spacetimes for quite general 2-surfaces not quite viable) generalization of the standard round sphere expression ( 27 ). This view appears to be supported by the fact that the Hawking energy has strange properties for very non-spherical surfaces, e. g. for 2-surfaces in Minkowski spacetime which are not metric spheres.14.2 On the Penrose mass
Penrose’s suggestion for the quasi-local mass (or, more generally, energy-momentum and angular momentum) was based on a promising and far-reaching strategy to use twistors at the fundamental level. The basic object of the construction, the so-called kinematical twistor, is intended to comprise both the energy-momentum and angular momentum, and is a well defined quasi-local quantity on generic spacelike surfaces homeomorphic to . It can be interpreted as the value of a quasi-local Hamiltonian, and the four independent 2-surface twistors play the role of the quasi-translations and quasi-rotations. The kinematical twistor was calculated for a large class of special 2-surfaces and gave acceptable results. However, the construction is not complete. First, the construction does not work for 2-surfaces whose topology is different from , and does not work even for certain topological 2-spheres for which the 2-surface twistor equation admits more than four independent solutions (‘exceptional 2-surfaces’). Second, two additional objects, the so-called infinity twistor and a Hermitian inner product on the space of 2-surface twistors, are needed to get the energy-momentum and angular momentum from the kinematical twistor and to ensure their reality. The latter is needed if we want to define the quasi-local mass as a norm of the kinematical twistor. However, no natural infinity twistor has been found, and no natural Hermitian scalar product can exist if the 2-surface cannot be embedded into a conformally flat spacetime. In addition, in the small surface calculations the quasi-local mass may be complex. If, however, we do not want to form invariants of the kinematical twistor (e. g. the mass), but we want to extract the energy-momentum and angular momentum from the kinematical twistor and we want them to be real, then only a special combination of the infinity twistor and the Hermitian scalar product, the so-called ‘bar-hook combination’ (see equation ( 48 )), would be needed. To save the main body of the construction the definition of the kinematical twistor was modified. Nevertheless, the mass in the modified constructions encountered an inherent ambiguity in the small surface approximation. One can still hope to find an appropriate ‘bar-hook’, and hence real energy-momentum and angular momentum, but invariants, such as norms, could not be formed.14.3 On the Dougan–Mason energy-momenta and the holomorphic/anti-holomorphic spin angular momenta
From pragmatic points of view the Dougan–Mason energy-momenta (subsection 8.2 ) are certainly among the most successful definitions: The energy-positivity and the rigidity (zero energy implies flatness), and the intimate connection between the pp-waves and the vanishing of the masses make these definitions potentially as useful quasi-local tools as the ADM and Bondi–Sachs energy-momenta in the asymptotically flat context. Similar properties are proven for the quasi-local energy-momentum of the matter fields, too. They depend only on the 2-surface data on , they have a clear Lagrangian interpretation, and the spinor fields that they are based on can be considered as the spinor constituents of the quasi-translations of the 2-surface. In fact, in the Minkowski spacetime the corresponding spacetime vectors are precisely the restriction to of the constant Killing vectors. These notions of energy-momentum are linked completely to the geometry of , and are independent of any ad hoc choice for the ‘fleet of observers’ on it. On the other hand, the holomorphic/anti-holomorphic spinor fields determine a six real parameter family of orthonormal frame fields on , which can be interpreted as some distinguished class of observers. In addition, they reproduce the expected, correct limits in a number of special situations. In particular, these energy-momenta appear to have been completed by spin-angular momenta (subsection 9.2 ) in a natural way. However, in spite of their successes, the Dougan–Mason energy-momenta and the spin-angular momenta based on Bramson’s superpotential and the holomorphic/anti-holomorphic spinor fields have some unsatisfactory properties, too (see the lists of our expectations in subsection 4.3 ). First, they are defined only for topological 2-spheres (but not for other topologies, e. g. for the torus ), and they are not well defined even for certain topological 2-spheres either. Such surfaces are, for example, past marginally trapped surfaces in the anti-holomorphic (and future marginally trapped surfaces in the holomorphic) case. Although the quasi-local mass associated with a marginally trapped surface is expected to be its irreducible mass , neither of the Dougan–Mason masses is well defined for the bifurcation surfaces of the Kerr–Newman (or even Schwarzschild) black hole. Second, the role and the physical content of the holomorphicity/anti-holomorphicity of the spinor fields is not clear. The use of the complex structure is justified a posteriori by the nice physical properties of the constructions and the pure mathematical fact that it is only the holomorphy and anti-holomorphy operators in a large class of potentially acceptable first order linear differential operators acting on spinor fields that have two dimensional kernel. Furthermore, since the holomorphic and anti-holomorphic constructions are not equivalent, we have two constructions instead of one, and it is not clear why we should prefer for example holomorphicity instead of anti-holomorphicity even at the quasi-local level. The angular momentum based on Bramson’s superpotential and the anti-holomorphic spinors together with the anti-holomorphic Dougan–Mason energy-momentum give acceptable Pauli–Lubanski spin for axis-symmetric zero-mass Cauchy developments, for small spheres, and at future null infinity, but the global angular momentum at the future null infinity is finite and well defined only if the spatial 3-momentum part of the Bondi–Sachs 4-momentum is vanishing, i. e. only in the centre-of-mass frame. (The spatial infinity limit of the spin-angular momenta has not been calculated.) Thus the Nester–Witten 2-form appears to serve as an appropriate framework for defining the energy-momentum, and it is the two spinor fields which should probably be changed and a new choice would be needed. The holomorphic/anti-holomorphic spinor fields appears to be ‘too rigid’. In fact, it is the topology of , namely the zero genus of , that restricts the solution space to two complex dimensions, instead of the local properties of the differential equations. (Thus, the situation is the same as in the twistorial construction of Penrose.) On the other hand, Bramson’s superpotential is based on the idea of Bergmann and Thomson that the angular momentum of gravity is analogous to the spin. Thus the question arises as to whether this picture is correct, or the gravitational angular momentum also has an orbital part, whenever Bramson’s superpotential describes only (the general form of) its spin part. The fact that our anti-holomorphic construction gives the correct, expected results for small spheres but unacceptable ones for large spheres near future null infinity in frames that are not centre-of-mass frames may indicate the lack of such an orbital term. This term could be neglected for small spheres, but certainly not for large spheres. For example, in the special quasi-local angular momentum of Bergqvist and Ludvigsen for the Kerr spacetime (subsection 9.3 ) it is the sum of Bramson’s expression and a term that can be interpreted as the orbital angular momentum.14.4 On the Brown–York type expressions
The idea of Brown and York that the quasi-local conserved quantities should be introduced via the canonical formulation of the theory is quite natural. In fact, as we saw, one could arrive at their general formulae from different points of departure (functional differentiability of the Hamiltonian, 2-surface observables). If the a priori requirement that we should have a well defined action principle for the trace- -action yielded undoubtedly well behaving quasi-local expressions, then the results would a posteriori justify this basic requirement (like the holomorphicity or anti-holomorphicity of the spinor fields in the Dougan–Mason definitions). However, if not, then that might be considered as an unnecessarily restrictive assumption and the question arises whether the present framework is wide enough to construct reasonable quasi-local energy-momentum and angular momentum. Indeed, the basic requirement automatically yields the boundary condition that the 3-metric should be fixed on the boundary , and that the boundary term in the Hamiltonian should be built only from the surface stress tensor . Since the boundary conditions are given, no Legendre transformation of the canonical variables on the 2-surface is allowed (see the derivation of Kijowski’s expression in subsection 10.2 ). The use of has important consequences. First, the quasi-local quantities depend not only on the geometry of the 2-surface , but on an arbitrarily chosen boost gauge, interpreted as a ‘fleet of observers being at rest with respect to ’, too. This leaves a huge ambiguity in the Brown–York energy (three functions of two variables, corresponding to the three boost parameters at each point of ) unless a natural gauge choice is prescribed. Second, since does not contain the extrinsic curvature of in the direction , which is a part of the 2-surface data, this extrinsic curvature is ‘lost’ from the point of view of the quasi-local quantities. Moreover, since is a tensor only on the 3-manifold , the integral of on is not sensitive to the component of normal to . The normal piece of the generator is ‘lost’ from the point of view of the quasi-local quantities. The other important ingredient of the Brown–York construction is the prescription of the subtraction term. Considering the Gauss–Codazzi–Mainardi equations of the isometric embedding of the 2-surface into the flat 3-space (or rather into a spacelike hyperplane of Minkowski spacetime) only as a system of differential equations for the reference extrinsic curvature, this prescription – contrary to frequently appearing opinions – is as explicit as the condition of the holomorphicity/anti-holomorphicity of the spinor fields in the Dougan–Mason definition. (One essential, and from pragmatic points of view important, difference is that the Gauss–Codazzi–Mainardi equations form an underdetermined elliptic system constrained by a nonlinear algebraic equation.) Similarly to the Dougan–Mason definitions, the general Brown–York formulae are valid for arbitrary spacelike 2-surfaces, but solutions to the equations defining the reference configuration exist certainly only for topological 2-spheres with strictly positive intrinsic scalar curvature. Thus there are exceptional 2-surfaces here, too. On the other hand, the Brown–York expressions (both for the flat 3-space and the light cone references) work properly for large spheres. At first sight, this choice for the definition of the subtraction term seems quite natural. However, we do not share this view. If the physical spacetime is the Minkowski one, then we expect that the geometry of the 2-surface in the reference Minkowski spacetime be the same as in the physical Minkowski spacetime. In particular, if – in the physical Minkowski spacetime – does not lie in any spacelike hyperplane, then we think that it would be un-natural to require the embedding of into a hyperplane of the reference Minkowski spacetime. Since in the two Minkowski spacetimes the extrinsic curvatures can be quite different, the quasi-local energy expressions based on this prescription of the reference term can be expected to yield nonzero value even in flat spacetime. Indeed, there are explicit examples showing this defect. (Epp’s definition is free of this difficulty, because he embeds the 2-surface into the Minkowski spacetime by preserving its ‘universal structure’. See subsection 4.1.4 .) Another objection against the embedding into flat 3-space is that it is not Lorentz covariant. As we discussed in subsection 4.2.2 , any spacetime covariant quasi-local energy expression for small spheres in vacuum should be of order with the Bel–Robinson ‘energy’ as the factor of proportionality. The Brown–York expression (even with the light cone reference ) fails to give the Bel–Robinson ‘energy’. Finally, in contrast to the Dougan–Mason definitions, the Brown–York type expressions are well defined on marginally trapped surfaces. However, they yield just twice the expected irreducible mass, and they do not reproduce the standard round sphere expression, which, for non-trapped surfaces, comes out from all the other expressions discussed in the present section (including Kijowski’s definition). It is remarkable that the derivation of the first law of black hole thermodynamics, based on the identification of the thermodynamical internal energy with the Brown–York energy, is independent of the definition of the subtraction term.It could be interesting to clarify the consequences of the boost gauge choice that is based on the main extrinsic curvature vector , discussed in subsection 4.1.2 . This would rule out the arbitrary element of the construction.
It might be interesting to see the small sphere expansion of the Kijowski and Kijowski–Liu–Yau expressions in vacuum.
15 Acknowledgments
I am grateful to Peter Aichelburg, Herbert Balasin, Robert Bartnik, Robert Beig, Piotr Chrusciel, Sergio Dain, Jorg Frauendiener, Sean Hayward, Jacek Jezierski, Jerzy Kijowski, Stephen Lau, Lionel Mason, Niall O Murchadha, James Nester, Ezra Newman, Alexander Petrov, Walter Simon, George Sparling, Paul Tod and Helmuth Urbantke for their valuable comments, remarks and stimulating questions. Special thanks to Jorg Frauendiener for continuous and fruitful discussions in the last eight years; to James Nester for the critical reading of an earlier version of the present manuscript, whose notes and remarks considerably improved its clarity; and to the two referees whose constructive criticism helped to make the present review more accurate and complete. Thanks are due to the Erwin Schrodinger Institut, Vienna, the Stefan Banach Center, Warsaw, the Universitat Tubingen, the Max Planck Institut fur Mathematik in den Naturwissenschaften, Leipzig, the National Center for Theoretical Sciences, Hsinchu, and the National Central University, Chungli, for hospitality, where parts of the present work were done and/or could be presented. This work was partially supported by the Hungarian Scientific Research Fund grants OTKA T030374 and T042531. References
Note: The reference version of this article is published by Living Reviews in Relativity