Open main menu
Home
Random
Recent changes
Special pages
Community portal
Preferences
About Wikipedia
Disclaimers
Incubator escapee wiki
Search
User menu
Talk
Dark mode
Contributions
Create account
Log in
Editing
General relativity
(section)
Warning:
You are not logged in. Your IP address will be publicly visible if you make any edits. If you
log in
or
create an account
, your edits will be attributed to your username, along with other benefits.
Anti-spam check. Do
not
fill this in!
== Advanced concepts == === Asymptotic symmetries === {{Main|BondiâMetznerâSachs group}} The spacetime symmetry group for [[special relativity]] is the [[PoincarĂ© group]], which is a ten-dimensional group of three Lorentz boosts, three rotations, and four spacetime translations. It is logical to ask what symmetries, if any, might apply in General Relativity. A tractable case might be to consider the symmetries of spacetime as seen by observers located far away from all sources of the gravitational field. The naive expectation for asymptotically flat spacetime symmetries might be simply to extend and reproduce the symmetries of flat spacetime of special relativity, ''viz.'', the PoincarĂ© group. In 1962 [[Hermann Bondi]], M. G. van der Burg, A. W. Metzner<ref name="bondi etal 1962">{{cite journal|title=Gravitational waves in general relativity: VII. Waves from axisymmetric isolated systems|journal= Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences|volume=269|pages=21â52|doi=10.1098/rspa.1962.0161|year=1962|last1=Bondi|first1=H.|last2=Van der Burg|first2=M.G.J.|last3=Metzner|first3=A.|issue=1336|bibcode=1962RSPSA.269...21B|s2cid=120125096}}</ref> and [[Rainer K. Sachs]]<ref name=sachs1962>{{cite journal|title=Asymptotic symmetries in gravitational theory|journal=Physical Review|volume=128|pages=2851â2864|doi=10.1103/PhysRev.128.2851|year=1962|last1=Sachs|first1=R.|issue=6|bibcode=1962PhRv..128.2851S}}</ref> addressed this [[BondiâMetznerâSachs group|asymptotic symmetry]] problem in order to investigate the flow of energy at infinity due to propagating [[gravitational wave]]s. Their first step was to decide on some physically sensible boundary conditions to place on the gravitational field at light-like infinity to characterize what it means to say a metric is asymptotically flat, making no ''a priori'' assumptions about the nature of the asymptotic symmetry groupânot even the assumption that such a group exists. Then after designing what they considered to be the most sensible boundary conditions, they investigated the nature of the resulting asymptotic symmetry transformations that leave invariant the form of the boundary conditions appropriate for asymptotically flat gravitational fields. What they found was that the asymptotic symmetry transformations actually do form a group and the structure of this group does not depend on the particular gravitational field that happens to be present. This means that, as expected, one can separate the kinematics of spacetime from the dynamics of the gravitational field at least at spatial infinity. The puzzling surprise in 1962 was their discovery of a rich infinite-dimensional group (the so-called BMS group) as the asymptotic symmetry group, instead of the finite-dimensional PoincarĂ© group, which is a subgroup of the BMS group. Not only are the Lorentz transformations asymptotic symmetry transformations, there are also additional transformations that are not Lorentz transformations but are asymptotic symmetry transformations. In fact, they found an additional infinity of transformation generators known as ''supertranslations''. This implies the conclusion that General Relativity (GR) does ''not'' reduce to special relativity in the case of weak fields at long distances. It turns out that the BMS symmetry, suitably modified, could be seen as a restatement of the universal [[soft graviton theorem]] in [[quantum field theory]] (QFT), which relates universal infrared (soft) QFT with GR asymptotic spacetime symmetries.<ref name=strominger2017>{{cite arXiv|title=Lectures on the Infrared Structure of Gravity and Gauge Theory|eprint=1703.05448|year=2017|last1=Strominger|first1=Andrew|class=hep-th|quote=...redacted transcript of a course given by the author at Harvard in spring semester 2016. It contains a pedagogical overview of recent developments connecting the subjects of soft theorems, the memory effect and asymptotic symmetries in four-dimensional QED, nonabelian gauge theory and gravity with applications to black holes. To be published Princeton University Press, 158 pages.}}</ref> === Causal structure and global geometry === {{Main|Causal structure}} [[File:Penrose.svg|thumb|PenroseâCarter diagram of an infinite [[Minkowski space|Minkowski universe]]]] In general relativity, no material body can catch up with or overtake a light pulse. No influence from an event ''A'' can reach any other location ''X'' before light sent out at ''A'' to ''X''. In consequence, an exploration of all light worldlines ([[Geodesics in general relativity|null geodesics]]) yields key information about the spacetime's causal structure. This structure can be displayed using [[Penrose diagram|PenroseâCarter diagrams]] in which infinitely large regions of space and infinite time intervals are shrunk ("[[Compactification (mathematics)|compactified]]") so as to fit onto a finite map, while light still travels along diagonals as in standard [[spacetime diagram]]s.<ref>{{Harvnb|Frauendiener|2004}}, {{Harvnb|Wald|1984|loc=sec. 11.1}}, {{Harvnb|Hawking|Ellis|1973|loc=sec. 6.8, 6.9}}</ref> Aware of the importance of causal structure, [[Roger Penrose]] and others developed what is known as [[Spacetime topology|global geometry]]. In global geometry, the object of study is not one particular solution (or family of solutions) to Einstein's equations. Rather, relations that hold true for all geodesics, such as the [[Raychaudhuri equation]], and additional non-specific assumptions about the nature of matter (usually in the form of [[energy conditions]]) are used to derive general results.<ref>{{Harvnb|Wald|1984|loc=sec. 9.2â9.4}} and {{Harvnb|Hawking|Ellis|1973|loc=ch. 6}}</ref> === Horizons === {{Main|Horizon (general relativity)|No hair theorem|Black hole mechanics}} Using global geometry, some spacetimes can be shown to contain boundaries called [[event horizon|horizons]], which demarcate one region from the rest of spacetime. The best-known examples are black holes: if mass is compressed into a sufficiently compact region of space (as specified in the [[hoop conjecture]], the relevant length scale is the [[Schwarzschild radius]]<ref>{{Harvnb|Thorne|1972}}; for more recent numerical studies, see {{Harvnb|Berger|2002|loc=sec. 2.1}}</ref>), no light from inside can escape to the outside. Since no object can overtake a light pulse, all interior matter is imprisoned as well. Passage from the exterior to the interior is still possible, showing that the boundary, the black hole's ''horizon'', is not a physical barrier.<ref>{{Harvnb|Israel|1987}}. A more exact mathematical description distinguishes several kinds of horizon, notably event horizons and [[apparent horizon]]s cf. {{Harvnb|Hawking|Ellis|1973|pp=312â320}} or {{Harvnb|Wald|1984|loc=sec. 12.2}}; there are also more intuitive definitions for isolated systems that do not require knowledge of spacetime properties at infinity, cf. {{Harvnb|Ashtekar|Krishnan|2004}}</ref> [[File:Ergosphere of a rotating black hole.svg|thumb|The ergosphere of a rotating black hole, which plays a key role when it comes to extracting energy from such a black hole]] Early studies of black holes relied on explicit solutions of Einstein's equations, notably the spherically symmetric Schwarzschild solution (used to describe a [[Static spacetime|static]] black hole) and the axisymmetric [[Kerr solution]] (used to describe a rotating, [[Stationary spacetime|stationary]] black hole, and introducing interesting features such as the ergosphere). Using global geometry, later studies have revealed more general properties of black holes. With time they become rather simple objects characterized by eleven parameters specifying: electric charge, massâenergy, [[linear momentum]], [[angular momentum]], and location at a specified time. This is stated by the [[no hair theorem|black hole uniqueness theorem]]: "black holes have no hair", that is, no distinguishing marks like the hairstyles of humans. Irrespective of the complexity of a gravitating object collapsing to form a black hole, the object that results (having emitted gravitational waves) is very simple.<ref>For first steps, cf. {{Harvnb|Israel|1971}}; see {{Harvnb|Hawking|Ellis|1973|loc=sec. 9.3}} or {{Harvnb|Heusler|1996|loc=ch. 9 and 10}} for a derivation, and {{Harvnb|Heusler|1998}} as well as {{Harvnb|Beig|ChruĆciel|2006}} as overviews of more recent results</ref> Even more remarkably, there is a general set of laws known as [[black hole mechanics]], which is analogous to the [[laws of thermodynamics]]. For instance, by the second law of black hole mechanics, the area of the event horizon of a general black hole will never decrease with time, analogous to the [[entropy]] of a thermodynamic system. This limits the energy that can be extracted by classical means from a rotating black hole (e.g. by the [[Penrose process]]).<ref>The laws of black hole mechanics were first described in {{Harvnb|Bardeen|Carter|Hawking|1973}}; a more pedagogical presentation can be found in {{Harvnb|Carter|1979}}; for a more recent review, see {{Harvnb|Wald|2001|loc=ch. 2}}. A thorough, book-length introduction including an introduction to the necessary mathematics {{Harvnb|Poisson|2004}}. For the Penrose process, see {{Harvnb|Penrose|1969}}</ref> There is strong evidence that the laws of black hole mechanics are, in fact, a subset of the laws of thermodynamics, and that the black hole area is proportional to its entropy.<ref>{{Harvnb|Bekenstein|1973}}, {{Harvnb|Bekenstein|1974}}</ref> This leads to a modification of the original laws of black hole mechanics: for instance, as the second law of black hole mechanics becomes part of the second law of thermodynamics, it is possible for the black hole area to decrease as long as other processes ensure that entropy increases overall. As thermodynamical objects with nonzero temperature, black holes should emit [[thermal radiation]]. Semiclassical calculations indicate that indeed they do, with the surface gravity playing the role of temperature in [[Planck's law]]. This radiation is known as [[Hawking radiation]] (cf. the [[#Quantum field theory in curved spacetime|quantum theory section]], below).<ref>The fact that black holes radiate, quantum mechanically, was first derived in {{Harvnb|Hawking|1975}}; a more thorough derivation can be found in {{Harvnb|Wald|1975}}. A review is given in {{Harvnb|Wald|2001|loc=ch. 3}}</ref> There are many other types of horizons. In an expanding universe, an observer may find that some regions of the past cannot be observed ("[[particle horizon]]"), and some regions of the future cannot be influenced (event horizon).<ref>{{Harvnb|Narlikar|1993|loc=sec. 4.4.4, 4.4.5}}</ref> Even in flat Minkowski space, when described by an accelerated observer ([[Rindler space]]), there will be horizons associated with a semiclassical radiation known as [[Unruh effect|Unruh radiation]].<ref>Horizons: cf. {{Harvnb|Rindler|2001|loc=sec. 12.4}}. Unruh effect: {{Harvnb|Unruh|1976}}, cf. {{Harvnb|Wald|2001|loc=ch. 3}}</ref> === Singularities === {{Main|Spacetime singularity}} Another general feature of general relativity is the appearance of spacetime boundaries known as singularities. Spacetime can be explored by following up on timelike and lightlike geodesicsâall possible ways that light and particles in free fall can travel. But some solutions of Einstein's equations have "ragged edges"âregions known as [[spacetime singularity|spacetime singularities]], where the paths of light and falling particles come to an abrupt end, and geometry becomes ill-defined. In the more interesting cases, these are "curvature singularities", where geometrical quantities characterizing spacetime curvature, such as the [[Ricci scalar]], take on infinite values.<ref>{{Harvnb|Hawking|Ellis|1973|loc=sec. 8.1}}, {{Harvnb|Wald|1984|loc=sec. 9.1}}</ref> Well-known examples of spacetimes with future singularitiesâwhere worldlines endâare the Schwarzschild solution, which describes a singularity inside an eternal static black hole,<ref>{{Harvnb|Townsend|1997|loc=ch. 2}}; a more extensive treatment of this solution can be found in {{Harvnb|Chandrasekhar|1983|loc=ch. 3}}</ref> or the Kerr solution with its ring-shaped singularity inside an eternal rotating black hole.<ref>{{Harvnb|Townsend|1997|loc=ch. 4}}; for a more extensive treatment, cf. {{Harvnb|Chandrasekhar|1983|loc=ch. 6}}</ref> The FriedmannâLemaĂźtreâRobertsonâWalker solutions and other spacetimes describing universes have past singularities on which worldlines begin, namely Big Bang singularities, and some have future singularities ([[Big Crunch]]) as well.<ref>{{Harvnb|Ellis|Van Elst|1999}}; a closer look at the singularity itself is taken in {{Harvnb|Börner|1993|loc=sec. 1.2}}</ref> Given that these examples are all highly symmetricâand thus simplifiedâit is tempting to conclude that the occurrence of singularities is an artifact of idealization.<ref>Here one should remind to the well-known fact that the important "quasi-optical" singularities of the so-called [[eikonal approximation]]s of many wave equations, namely the "[[caustic (mathematics)|caustics]]", are resolved into finite peaks beyond that approximation.</ref> The famous [[singularity theorems]], proved using the methods of global geometry, say otherwise: singularities are a generic feature of general relativity, and unavoidable once the collapse of an object with realistic matter properties has proceeded beyond a certain stage<ref>Namely when there are [[trapped null surface]]s, cf. {{Harvnb|Penrose|1965}}</ref> and also at the beginning of a wide class of expanding universes.<ref>{{Harvnb|Hawking|1966}}</ref> However, the theorems say little about the properties of singularities, and much of current research is devoted to characterizing these entities' generic structure (hypothesized e.g. by the [[BKL singularity|BKL conjecture]]).<ref>The conjecture was made in {{Harvnb|Belinskii|Khalatnikov|Lifschitz|1971}}; for a more recent review, see {{Harvnb|Berger|2002}}. An accessible exposition is given by {{Harvnb|Garfinkle|2007}}</ref> The [[cosmic censorship hypothesis]] states that all realistic future singularities (no perfect symmetries, matter with realistic properties) are safely hidden away behind a horizon, and thus invisible to all distant observers. While no formal proof yet exists, numerical simulations offer supporting evidence of its validity.<ref>The restriction to future singularities naturally excludes initial singularities such as the big bang singularity, which in principle be visible to observers at later cosmic time. The cosmic censorship conjecture was first presented in {{Harvnb|Penrose|1969}}; a textbook-level account is given in {{Harvnb|Wald|1984|pp=302â305}}. For numerical results, see the review {{Harvnb|Berger|2002|loc=sec. 2.1}}</ref> === Evolution equations === {{Main|Initial value formulation (general relativity)}} Each solution of Einstein's equation encompasses the whole history of a universeâit is not just some snapshot of how things are, but a whole, possibly matter-filled, spacetime. It describes the state of matter and geometry everywhere and at every moment in that particular universe. Due to its general covariance, Einstein's theory is not sufficient by itself to determine the [[time evolution]] of the metric tensor. It must be combined with a [[coordinate condition]], which is analogous to [[gauge fixing]] in other field theories.<ref>{{Harvnb|Hawking|Ellis|1973|loc=sec. 7.1}}</ref> To understand Einstein's equations as partial differential equations, it is helpful to formulate them in a way that describes the evolution of the universe over time. This is done in "3+1" formulations, where spacetime is split into three space dimensions and one time dimension. The best-known example is the [[ADM formalism]].<ref>{{Harvnb|Arnowitt|Deser|Misner|1962}}; for a pedagogical introduction, see {{Harvnb|Misner|Thorne|Wheeler|1973|loc=§ 21.4â§ 21.7}}</ref> These decompositions show that the spacetime evolution equations of general relativity are well-behaved: solutions always [[existence theorem|exist]], and are uniquely defined, once suitable initial conditions have been specified.<ref>{{Harvnb|FourĂšs-Bruhat|1952}} and {{Harvnb|Bruhat|1962}}; for a pedagogical introduction, see {{Harvnb|Wald|1984|loc=ch. 10}}; an online review can be found in {{Harvnb|Reula|1998}}</ref> Such formulations of Einstein's field equations are the basis of numerical relativity.<ref>{{Harvnb|Gourgoulhon|2007}}; for a review of the basics of numerical relativity, including the problems arising from the peculiarities of Einstein's equations, see {{Harvnb|Lehner|2001}}</ref> === Global and quasi-local quantities === {{Main|Mass in general relativity}} The notion of evolution equations is intimately tied in with another aspect of general relativistic physics. In Einstein's theory, it turns out to be impossible to find a general definition for a seemingly simple property such as a system's total mass (or energy). The main reason is that the gravitational fieldâlike any physical fieldâmust be ascribed a certain energy, but that it proves to be fundamentally impossible to localize that energy.<ref>{{Harvnb|Misner|Thorne|Wheeler|1973|loc=§ 20.4}}</ref> Nevertheless, there are possibilities to define a system's total mass, either using a hypothetical "infinitely distant observer" ([[ADM mass]])<ref>{{Harvnb|Arnowitt|Deser|Misner|1962}}</ref> or suitable symmetries ([[Komar mass]]).<ref>{{Harvnb|Komar|1959}}; for a pedagogical introduction, see {{Harvnb|Wald|1984|loc=sec. 11.2}}; although defined in a totally different way, it can be shown to be equivalent to the ADM mass for stationary spacetimes, cf. {{Harvnb|Ashtekar|Magnon-Ashtekar|1979}}</ref> If one excludes from the system's total mass the energy being carried away to infinity by gravitational waves, the result is the [[Mass in general relativity#ADM and Bondi masses in asymptotically flat space-times|Bondi mass]] at null infinity.<ref>For a pedagogical introduction, see {{Harvnb|Wald|1984|loc=sec. 11.2}}</ref> Just as in [[Physics in the Classical Limit|classical physics]], it can be shown that these masses are positive.<ref>{{Harvnb|Wald|1984|p=295 and refs therein}}; this is important for questions of stabilityâif there were [[negative mass]] states, then flat, empty Minkowski space, which has mass zero, could evolve into these states</ref> Corresponding global definitions exist for momentum and angular momentum.<ref>{{Harvnb|Townsend|1997|loc=ch. 5}}</ref> There have also been a number of attempts to define ''quasi-local'' quantities, such as the mass of an isolated system formulated using only quantities defined within a finite region of space containing that system. The hope is to obtain a quantity useful for general statements about [[isolated system]]s, such as a more precise formulation of the hoop conjecture.<ref>Such quasi-local massâenergy definitions are the [[Hawking energy]], [[Geroch energy]], or Penrose's quasi-local energyâmomentum based on [[Twistor theory|twistor]] methods; cf. the review article {{Harvnb|Szabados|2004}}</ref>
Edit summary
(Briefly describe your changes)
By publishing changes, you agree to the
Terms of Use
, and you irrevocably agree to release your contribution under the
CC BY-SA 4.0 License
and the
GFDL
. You agree that a hyperlink or URL is sufficient attribution under the Creative Commons license.
Cancel
Editing help
(opens in new window)