Chapter 1Free to read

General relativity essentials


Published Copyright © IOP Publishing Ltd 2017
Pages 1-1 to 1-33

Download ePub chapter

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

Download complete PDF book, the ePub book or the Kindle book

Export citation and abstract

BibTeX RIS

Share this chapter

978-0-7503-1478-7

Abstract

The goal in this first chapter is to go in an efficient way through the essential material of general relativity and get quickly to Einstein's equations. A classic text on general relativity is Wald and a much newer text which has become a classic in its own right is Carroll.

This article is available under the terms of the IOP-Standard Books License

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means, electronic, mechanical, photocopying, recording or otherwise, without the prior permission of the publisher, or as expressly permitted by law or under terms agreed with the appropriate rights organization. Multiple copying is permitted in accordance with the terms of licences issued by the Copyright Licensing Agency, the Copyright Clearance Centre and other reproduction rights organisations.

Permission to make use of IOP Publishing content other than as set out above may be sought at permissions@iop.org.

This ebook contains interactive Q&A technology, allowing the reader to interact with the text and reveal answers to selected exercises posed by the author within the book. This feature may not function in all formats and on reading devices.

Certain images in this publication have been obtained by the author from the Wikipedia/Wikimedia website, where they were made available under a Creative Commons licence or stated to be in the public domain. Please see individual figure captions in this publication for details. To the extent that the law allows, IOP Publishing disclaim any liability that any person may suffer as a result of accessing, using or forwarding the image(s). Any reuse rights should be checked and permission should be sought if necessary from Wikipedia/Wikimedia and/or the copyright owner (as appropriate) before using or forwarding the image(s).

Badis Ydri has asserted his right to be identified as the author of this work in accordance with sections 77 and 78 of the Copyright, Designs and Patents Act 1988.

The goal in this first chapter is to cover the essential material of general relativity in an efficient manner, and get quickly to Einstein equations. A classic text on general relativity is by Wald [1] and a much newer text which has become a classic in its own right is by Carroll [2]. We have also drawn from t'Hooft [3] and from many other texts and lecture notes on the subject, which are cited in subsequent chapters.

1.1. The equivalence principle

The classical (Newtonian) theory of gravity is based on the following two equations. The gravitational potential Φ generated by a mass density ρ is given by Poisson's equations (with G being Newton's constant)

Equation (1.1)

The force exerted by this potential Φ on a particle of mass m is given by

Equation (1.2)

These equations are obviously not compatible with the special theory of relativity. The above first equation will be replaced, in the general relativistic theory of gravity, by Einstein's equations of motion, while the second equation will be replaced by the geodesic equation. From the above two equations we see that there are two measures of gravity: ${{\rm{\nabla }}}^{2}{\rm{\Phi }}$ measures the source of gravity, while $\vec{{\rm{\nabla }}}{\rm{\Phi }}$ measures the effect of gravity. Thus $\vec{{\rm{\nabla }}}{\rm{\Phi }}$, outside a source of gravity where $\rho ={{\rm{\nabla }}}^{2}{\rm{\Phi }}=0$, need not vanish. The analogues of these two different measures of gravity, in general relativity, are given by the so-called Ricci curvature tensor ${R}_{\mu \nu }$ and Riemann curvature tensor ${R}_{\mu \nu \alpha \beta }$, respectively.

The basic postulate of general relativity is simply that gravity is geometry. More precisely, gravity will be identified with the curvature of spacetime which is taken to be a pseudo-Riemannian (Lorentzian) manifold. This can be made more precise by employing the two guiding 'principles' which led Einstein to his equations. These are:

  • •  
    The weak equivalence principle. This states that all particles fall the same way in a gravitational field which is equivalent to the fact that the inertial mass is identical to the gravitational mass. In other words, the dynamics of all free particles, falling in a gravitational field, is completely specified by a single worldline. This is to be contrasted with charged particles in an electric field which obviously follow different worldlines depending on their electric charges. Thus, at any point in spacetime, the effect of gravity is fully encoded in the set of all possible worldlines, corresponding to all initial velocities, passing at that point. These worldlines are precisely the so-called geodesics.In measuring the electromagnetic field we choose 'background observers' who are not subject to electromagnetic interactions. These are clearly inertial observers who follow geodesic motion. The worldline of a charged test body can then be measured by observing the deviation from the inertial motion of the observers.
  •    
    This procedure cannot be applied to measure the gravitational field since by the equivalence principle gravity acts the same way on all bodies, i.e. we cannot insulate the 'background observers' from the effect of gravity so that they provide inertial observers. In fact, any observer will move under the effect of gravity in exactly the same way as the test body.
  •    
    The central assumption of general relativity is that we cannot, even in principle, construct inertial observers who follow geodesic motion and measure the gravitational force. Indeed, we assume that the spacetime metric is curved and that the worldlines of freely falling bodies in a gravitational field are precisely the geodesics of the curved metric. In other words, the 'background observers' which are the geodesics of the curved metric coincide exactly with motion in a gravitational field.
  •    
    Therefore, gravity is not a force since it cannot be measured, but is a property of spacetime. Gravity is in fact the curvature of spacetime. The gravitational field corresponds thus to a deviation of the spacetime geometry from the flat geometry of special relativity. But infinitesimally each manifold is flat. This leads us to the Einstein's equivalence principle: in small enough regions of spacetime, the non-gravitational laws of physics reduce to special relativity since it is not possible to detect the existence of a gravitational field through local experiments.
  • •  
    Mach's principle. This states that all matter in the Universe must contribute to the local definition of 'inertial motion' and 'non-rotating motion'. Equivalently the concepts of 'inertial motion' and 'non-rotating motion' are meaningless in an empty Universe. In the theory of general relativity the distribution of matter in the Universe indeed influences the structure of spacetime. In contrast, the theory of special relativity asserts that 'inertial motion' and 'non-rotating motion' are not influenced by the distribution of matter in the Universe.

Therefore, in general relativity the laws of physics must:

  • (i)  
    reduce to the laws of physics in special relativity in the limit where the metric ${g}_{\mu \nu }$ becomes flat or in a sufficiently small region around a given point in spacetime.
  • (ii)  
    be covariant under general coordinate transformations, which generalizes the covariance under Poincaré found in special relativity. This means in particular that only the metric ${g}_{\mu \nu }$ and quantities derived from it can appear in the laws of physics.

In summary, general relativity is the theory of space, time, and gravity in which spacetime is a curved manifold M, which is not necessarily R4, on which a Lorentzian metric ${g}_{\mu \nu }$ is defined. The curvature of spacetime in this metric is related to the stress–energy–momentum tensor of the matter in the Universe, which is the source of gravity, by Einstein's equations which are schematically given by equations of the form

Equation (1.3)

This is the analogue of equation (1.1). The worldlines of freely falling objects in this gravitational field are precisely given by the geodesics of this curved metric. In small enough regions of spacetime, curvature vanishes, i.e. spacetime becomes flat, and the geodesic become straight. Thus, the analogue of equation (1.2) is given schematically by an equation of the form

Equation (1.4)

1.2. Relativistic mechanics

In special relativity spacetime has the manifold structure R4 with a flat metric of Lorentzian signature defined on it. In special relativity, as in pre-relativity physics, an inertial motion is one in which the observer or the test particle is non-accelerating, which obviously corresponds to no external forces acting on the observer or the test particle. An inertial observer at the origin of spacetime can construct a rigid frame where the grid points are labeled by ${x}^{1}=x$, ${x}^{2}=y$, and ${x}^{3}=z$. Furthermore, she/he can equip the grid points with synchronized clocks which give the reading ${x}^{0}={ct}$. This provides a global inertial coordinate system or reference frame of spacetime where every point is labeled by $({x}^{0},{x}^{1},{x}^{2},{x}^{3})$. The labels have no intrinsic meaning, but the interval between two events A and B defined by $-{\left({x}_{A}^{0}-{x}_{B}^{0}\right)}^{2}+{\left({x}_{A}^{i}-{x}_{B}^{i}\right)}^{2}$ is an intrinsic property of spacetime since its value is the same in all global inertial reference frames. The metric tensor of spacetime in a global inertial reference frame $\{{x}^{\mu }\}$ is a tensor of type $(0,2)$ with components ${\eta }_{\mu \nu }=(-1,+1,+1,+1)$, i.e. ${{ds}}^{2}=-{({{dx}}^{0})}^{2}+{({{dx}}^{i})}^{2}$. The derivative operator associated with this metric is the ordinary derivative, and as a consequence the curvature of this metric vanishes. The geodesics are straight lines. The time-like geodesics are precisely the world lines of inertial observables.

Let ta be the tangent of a given curve in spacetime. The norm ${\eta }_{\mu \nu }{t}^{\mu }{t}^{\nu }$ is positive, negative, and zero for space-like, time-like, and light-like (null) curves, respectively. Since material objects cannot travel faster than light their paths in spacetime must be time-like. The proper time along a time-like curve parameterized by t is defined by

Equation (1.5)

This proper time is the elapsed time on a clock carried on the time-like curve. The so-called 'twin paradox' is the statement that different time-like curves connecting two points have different proper times. The curve with maximum proper time is the geodesic connecting the two points in question. This curve corresponds to inertial motion between the two points.

The 4-vector velocity of a massive particle with a 4-vector position ${x}^{\mu }$ is ${U}^{\mu }={{dx}}^{\mu }/d\tau $ where τ is the proper time. Clearly we must have ${U}^{\mu }{U}_{\mu }=-{c}^{2}$. In general, the tangent vector ${U}^{\mu }$ of a time-like curve parameterized by the proper time τ will be called the 4-vector velocity of the curve and it will satisfy

Equation (1.6)

A free particle will be in inertial motion. The trajectory will therefore be given by a time-like geodesic given by the equation

Equation (1.7)

Indeed, the operator ${U}^{\mu }{\partial }_{\mu }$ is the directional derivative along the curve. The energy–momentum 4-vector ${p}^{\mu }$ of a particle with rest mass m is given by

Equation (1.8)

This leads to (with $\gamma =1/\sqrt{1-{\vec{u}}^{2}/{c}^{2}}$ and $\vec{u}=d\vec{x}/{dt}$)

Equation (1.9)

We also compute

Equation (1.10)

The energy of a particle as measured by an observer whose velocity is ${v}^{\mu }$ is then clearly given by

Equation (1.11)

1.3. Differential geometry primer

1.3.1. Metric manifolds and vectors

Metric manifolds. An n-dimensional manifold M is a space which is locally flat, i.e. locally looks like Rn , and furthermore can be constructed from pieces of Rn sewn together smoothly. A Lorentzian or pseudo-Riemannian manifold is a manifold with the notion of 'distance', equivalently 'metric', included. 'Lorentzian' refers to the signature of the metric which in general relativity is taken to be $(-1,+1,+1,+1)$ as opposed to the more familiar/natural 'Euclidean' signature given by $(+1,+1,+1,+1)$ valid for Riemannian manifolds. The metric is usually denoted by ${g}_{\mu \nu }$ while the line element (also called metric in many instances) is written as

Equation (1.12)

For example Minkowski spacetime is given by the flat metric

Equation (1.13)

Another extremely important example is Schwarzschild spacetime given by the metric

Equation (1.14)

This is quite different from the flat metric ${\eta }_{\mu \nu }$ and as a consequence the curvature of Schwarzschild spacetime is non-zero. Another important curved space is the surface of the two-dimensional sphere on which the metric, which appears as a part of the Schwarzschild metric, is given by

Equation (1.15)

The inverse metric will be denoted by ${g}^{\mu \nu }$, i.e.

Equation (1.16)

Charts. A coordinate system (a chart) on the manifold M is a subset U of M together with a one-to-one map $\phi :U\to {R}^{n}$ such that the image $V=\phi (U)$ is an open set in Rn , i.e. a set in which every point $y\in V$ is the center of an open ball which is inside V. We say that U is an open set in M. Hence we can associate with every point $p\in U$ of the manifold M the local coordinates $({x}^{1},\ldots ,{x}^{n})$ by

Equation (1.17)

Vectors. A curved manifold is not necessarily a vector space. For example the sphere is not a vector space because we do not know how to add two points on the sphere to get another point on the sphere. The sphere, which is naturally embedded in R3, admits at each point p a tangent plane. The notion of a 'tangent vector space' can be constructed for any manifold which is embedded in Rn . The tangent vector space at a point p of the manifold will be denoted by Vp .

There is a one-to-one correspondence between vectors and directional derivatives in Rn . Indeed, the vector $v=({v}^{1},\ldots ,{v}^{n})$ in Rn defines the directional derivative ${\sum }_{\mu }{v}^{\mu }{\partial }_{\mu }$ which acts on functions on Rn . These derivatives are clearly linear and satisfy the Leibniz rule. We will therefore define tangent vectors at a given point p on a manifold M as directional derivatives which satisfy linearity and the Leibniz rule. These directional derivatives can also be thought of as differential displacements on the spacetime manifold at the point p.

This can be made more precise as follows. First, we define s smooth curve on the manifold M as a smooth map from R into M, namely $\gamma :R\to M$. A tangent vector at a point p can then be thought of as a directional derivative operator along a curve which goes through p. Indeed, a tangent vector T at $p=\gamma (t)\in M$, acting on smooth functions f on the manifold M, can be defined by

Equation (1.18)

In a given chart ϕ the point p will be given by $p={\phi }^{-1}(x)$ where $x\;=({x}^{1},\ldots ,{x}^{n})\in {R}^{n}$. Hence $\gamma (t)={\phi }^{-1}(x)$. In other words, the map γ is mapped into a curve x(t) in Rn . We have immediately

Equation (1.19)

The maps ${X}_{\mu }$ act on functions f on the Manifold M as

Equation (1.20)

These can be checked to satisfy linearity and the Leibniz rule. They are obviously directional derivatives or differential displacements since we may make the identification ${X}_{\mu }={\partial }_{\mu }$. Hence these vectors are tangent vectors to the manifold M at p. The fact that arbitrary tangent vectors can be expressed as linear combinations of the n vectors ${X}_{\mu }$ shows that these vectors are linearly independent, span the vector space Vp and that the dimension of Vp is exactly n. Equation (1.19) can then be rewritten as

Equation (1.21)

The components ${T}^{\mu }$ of the vector T are therefore given by

Equation (1.22)

1.3.2. Geodesics

The length l of a smooth curve C with tangent ${T}^{\mu }$ on a manifold M with Riemannian metric ${g}_{\mu \nu }$ is given by

Equation (1.23)

The length is parametrization independent. Indeed, we can show that

Equation (1.24)

In a Lorentzian manifold, the length of a space-like curve is also given by this expression. For a time-like curve for which ${g}_{{ab}}{T}^{a}{T}^{b}\lt 0$ the length is replaced with the proper time τ, which is given by $\tau =\int {dt}\sqrt{-{g}_{{ab}}{T}^{a}{T}^{b}}$. For a light-like (or null) curve for which ${g}_{{ab}}{T}^{a}{T}^{b}=0$ the length is always 0.

We consider the length of a curve C connecting two points $p=C({t}_{0})$ and $q=C({t}_{1})$. In a coordinate basis the length is given explicitly by

Equation (1.25)

The variation in l under an arbitrary smooth deformation of the curve C which keeps the two points p and q fixed is given by

Equation (1.26)

We can assume without any loss of generality that the parametrization of the curve C satisfies ${g}_{\mu \nu }({{dx}}^{\mu }/{dt})({{dx}}^{\nu }/{dt})=1$. In other words, we choose dt2 to be precisely the line element (interval) and thus ${T}^{\mu }={{dx}}^{\mu }/{dt}$ is the 4-velocity. The last term in the above equation obviously becomes a total derivative, which vanishes by the fact that the considered deformation keeps the two end points p and q fixed. We then obtain

Equation (1.27)

By definition geodesics are curves which extremize the length l. The curve C extremizes the length between the two points p and q if and only if $\delta l=0$. This leads immediately to the equation

Equation (1.28)

This equation is called the geodesic equation. It is the relativistic generalization of Newton's second law of motion (1.2). The Christoffel symbols are defined by

Equation (1.29)

In the absence of curvature we will have ${g}_{\mu \nu }={\eta }_{\mu \nu }$ and hence ${\rm{\Gamma }}=0$. In other words, the geodesics are locally straight lines.

Since the length between any two points on a Riemannian manifold (and between any two points which can be connected by a space-like curve on a Lorentzian manifold) can be arbitrarily long, we conclude that the shortest curve connecting the two points must be a geodesic as it is an extremum of length. Hence the shortest curve is the straightest possible curve. The converse is not true: a geodesic connecting two points is not necessarily the shortest path.

Similarly, the proper time between any two points which can be connected by a time-like curve on a Lorentzian manifold can be arbitrarily small and thus the curve with the greatest proper time, if it exists, must be a time-like geodesic as it is an extremum of proper time. On the other hand, a time-like geodesic connecting two points is not necessarily the path with maximum proper time.

1.3.3. Tensors

Tangent (contravariant) vectors. Tensors are a generalization of vectors. Let us start then by giving a more precise definition of the tangent vector space Vp . Let ${ \mathcal F }$ be the set of all smooth functions f on the manifold M, i.e. $f:M\to R$. We define a tangent vector v at the point $p\in M$ as a map $v:{ \mathcal F }\to R$ which is required to satisfy linearity and the Leibniz rule. In other words,

Equation (1.30)

The vector space Vp is simply the set of all tangents vectors v at p. The action of the vector v on the function f is given explicitly by

Equation (1.31)

In a different chart $\phi ^{\prime} $ we will have

Equation (1.32)

We compute

Equation (1.33)

This is why the basis elements ${X}_{\mu }$ may be thought of as the partial derivative operators $\partial /\partial {x}^{\mu }$. The tangent vector v can be rewritten as $v={\sum }_{\mu =1}^{n}{v}^{\mu }{X}_{\mu }={\sum }_{\mu =1}^{n}{v}^{^{\prime} \mu }{X}_{\mu }^{^{\prime} }$. We conclude immediately that

Equation (1.34)

This is the transformation law of tangent vectors under the coordinate transformation ${x}^{\mu }\to {x}^{^{\prime} \mu }$.

Cotangent dual (covariant) vectors or 1-forms. Let ${V}_{p}^{*}$ be the space of all linear maps ${\omega }^{*}$ from Vp into R, namely ${\omega }^{*}:{V}_{p}\to R$. The space ${V}_{p}^{*}$ is the so-called dual vector space to Vp where addition and multiplication by scalars are defined in an obvious way. The elements of ${V}_{p}^{*}$ are called dual vectors. The dual vector space ${V}_{p}^{*}$ is also called the cotangent dual vector space at p and the vector space of one-forms at p. The elements of ${V}_{p}^{*}$ are then called cotangent dual vectors. Another nomenclature is to refer to the elements of ${V}_{p}^{*}$ as covariant vectors, as opposed to the elements of Vp which are referred to as contravariant vectors.

The basis $\{{X}^{\mu *}\}$ of ${V}_{p}^{*}$ is called the dual basis to the basis $\{{X}_{\mu }\}$ of Vp . The basis elements of ${V}_{p}^{*}$ are given by vectors ${X}^{\mu *}$ defined by

Equation (1.35)

We have the transformation law

Equation (1.36)

From this result we can think of the basis elements ${X}^{\mu *}$ as the gradients ${{dx}}^{\mu }$, namely

Equation (1.37)

Let $v={\sum }_{\mu }{v}^{\mu }{X}_{\mu }$ be an arbitrary tangent vector in Vp , then the action of the dual basis elements ${X}^{\mu *}$ on v is given by

Equation (1.38)

The action of a general element ${\omega }^{*}={\sum }_{\mu }{\omega }_{\mu }{X}^{\mu *}$ of ${V}_{p}^{*}$ on v is given by

Equation (1.39)

Again we conclude the transformation law

Equation (1.40)

Generalization. A tensor T of type (k, l) over the tangent vector space Vp is a multilinear map form $({V}_{p}^{*}\times {V}_{p}^{*}\times \cdots \; \times {V}_{p}^{*})\times ({V}_{p}\times {V}_{p}\times \cdots \; \times {V}_{p})$ into R given by

Equation (1.41)

The domain of this map is the direct product of k cotangent dual vector space ${V}_{p}^{*}$ and l tangent vector space Vp . The space ${ \mathcal T }\quad (k,l)$ of all tensors of type (k, l) is a vector space of dimension ${n}^{k}.{n}^{l}$ since $\mathrm{dim}{V}_{p}=\mathrm{dim}{V}_{p}^{*}=n$.

The tangent vectors $v\in {V}_{p}$ are therefore tensors of type $(1,0)$, whereas the cotangent dual vectors $v\in {V}_{p}^{*}$ are tensors of type $(0,1)$. The metric g is a tensor of type $(0,2)$, i.e. a linear map from into R, which is symmetric and non-degenerate.

1.4. Curvature tensor

1.4.1. Covariant derivative

A covariant derivative is a derivative which transforms covariantly under coordinates transformations $x\to x^{\prime} $. In other words, it is an operator ∇ on the manifold M which takes a differentiable tensor of type (k, l) to a differentiable tensor of type $(k,l+1)$. It must clearly satisfy the obvious properties of linearity and Leibniz rule but also satisfies other important rules such as the torsion free condition given by

Equation (1.42)

Furthermore, the covariant derivative acting on scalars must be consistent with tangent vectors being directional derivatives. Indeed, for all $f\in { \mathcal F }$ and ${t}^{\mu }\in {V}_{p}$ we must have

Equation (1.43)

In other words, If ∇and $\tilde{{\rm{\nabla }}}$ are two covariant derivative operators, then their action on scalar functions must coincide, i.e.

Equation (1.44)

We now compute the difference ${\tilde{{\rm{\nabla }}}}_{\mu }( f{\omega }_{\nu })-{{\rm{\nabla }}}_{\mu }( f{\omega }_{\nu })$ where ω is some cotangent dual vector. We have

Equation (1.45)

We use without proof the following result. Let ${\omega }_{\nu }^{^{\prime} }$ be the value of the cotangent dual vector ${\omega }_{\nu }$ at a nearby point $p^{\prime} $, i.e. ${\omega }_{\nu }^{^{\prime} }-{\omega }_{\nu }$ is zero at p. Since the cotangent dual vector ${\omega }_{\nu }$ is a smooth function on the manifold, then for each $p^{\prime} \in M$, there must exist smooth functions ${f}_{(\alpha )}$ which vanish at the point p and cotangent dual vectors ${\mu }_{\nu }^{(\alpha )}$ such that

Equation (1.46)

We compute immediately

Equation (1.47)

This is 0 since by assumption ${f}_{(\alpha )}$ vanishes at p. Hence we obtain the result

Equation (1.48)

In other words, the difference ${\tilde{{\rm{\nabla }}}}_{\mu }{\omega }_{\nu }-{{\rm{\nabla }}}_{\mu }{\omega }_{\nu }$ depends only on the value of ${\omega }_{\nu }$ at the point p, although both ${\tilde{{\rm{\nabla }}}}_{\mu }{\omega }_{\nu }$ and ${{\rm{\nabla }}}_{\mu }{\omega }_{\nu }$ depend on how ${\omega }_{\nu }$ changes as we go away from point p since they are derivatives. Putting this differently we say that the operator ${\tilde{{\rm{\nabla }}}}_{\mu }-{{\rm{\nabla }}}_{\mu }$ is a linear map which takes cotangent dual vectors at a point p into tensors, of type $(0,2)$, at p and not into tensor fields defined in a neighborhood of p. We write

Equation (1.49)

The tensor ${C}_{\mu \nu }^{\gamma }$ stands for the map ${\tilde{{\rm{\nabla }}}}_{\mu }-{{\rm{\nabla }}}_{\mu }$ and it is clearly a tensor of type $(1,2)$. By setting ${\omega }_{\mu }={{\rm{\nabla }}}_{\mu }f={\tilde{{\rm{\nabla }}}}_{\mu }f$ we obtain ${{\rm{\nabla }}}_{\mu }{{\rm{\nabla }}}_{\nu }f={\tilde{{\rm{\nabla }}}}_{\mu }{\tilde{{\rm{\nabla }}}}_{\nu }f-{C}_{\mu \nu }^{\gamma }{{\rm{\nabla }}}_{\gamma }f$. By employing now the torsion free condition (1.42) we immediately obtain

Equation (1.50)

Let us consider now the difference ${\tilde{{\rm{\nabla }}}}_{\mu }({\omega }_{\nu }{t}^{\nu })-{{\rm{\nabla }}}_{\mu }({\omega }_{\nu }{t}^{\nu })$ where ${t}^{\nu }$ is a tangent vector. Since ${\omega }_{\nu }{t}^{\nu }$ is a function we have

Equation (1.51)

On the other hand, we compute

Equation (1.52)

Hence, we must have

Equation (1.53)

For a general tensor ${T}^{{\mu }_{1}\ldots {\mu }_{k}}{}_{{\nu }_{1}\ldots {\nu }_{l}}$ of type (k, l) the action of the covariant derivative operator will be given by the expression

Equation (1.54)

1.4.2. Parallel transport

Let C be a curve with a tangent vector ${t}^{\mu }$. Let ${v}^{\mu }$ be some tangent vector defined at each point on the curve. The vector ${v}^{\mu }$ is parallel transported along the curve C if and only if

Equation (1.55)

If t is the parameter along the curve C then ${t}^{\mu }={{dx}}^{\mu }/{dt}$ are the components of the vector ${t}^{\mu }$ in the coordinate basis. The parallel transport condition reads explicitly

Equation (1.56)

By demanding that the inner product of two vectors ${v}^{\mu }$ and ${w}^{\mu }$ is invariant under parallel transport we obtain, for all curves and all vectors, the condition

Equation (1.57)

Thus given a metric ${g}_{\mu \nu }$ on a manifold M the most natural covariant derivative operator is the one under which the metric is covariantly constant.

There exists a unique covariant derivative operator ${{\rm{\nabla }}}_{\mu }$ which satisfies ${{\rm{\nabla }}}_{\mu }{g}_{\alpha \beta }=0$. The proof goes as follows. We know that ${{\rm{\nabla }}}_{\mu }{g}_{\alpha \beta }$ is given by

Equation (1.58)

By imposing ${{\rm{\nabla }}}_{\mu }{g}_{\alpha \beta }=0$ we obtain

Equation (1.59)

Equivalently

Equation (1.60)

Equation (1.61)

Immediately, we conclude that

Equation (1.62)

In other words,

Equation (1.63)

This choice of ${C}_{\mu \alpha }^{\gamma }$ which solves ${{\rm{\nabla }}}_{\mu }{g}_{\alpha \beta }=0$ is unique. In other words, the corresponding covariant derivative operator is unique. The most important case corresponds to the choice ${\tilde{{\rm{\nabla }}}}_{a}={\partial }_{a}$ for which case ${C}_{ab}^{c}$ is denoted ${{\rm{\Gamma }}}_{ab}^{c}$ and is called the Christoffel symbol.

Equation (1.56) is almost the geodesic equation. Recall that geodesics are the straightest possible lines on a curved manifold. Alternatively, a geodesic can be defined as a curve whose tangent vector ${t}^{\mu }$ is parallel transported along itself, i.e. ${t}^{\mu }{{\rm{\nabla }}}_{\mu }{t}^{\nu }=0$, This reads in a coordinate basis as

Equation (1.64)

This is precisely equation (1.28). This is a set of n-coupled second order ordinary differential equations with n unknown ${x}^{\mu }(t)$. We know, given appropriate initial conditions ${x}^{\mu }({t}_{0})$ and ${{dx}}^{\mu }/{dt}{| }_{t={t}_{0}}$, that there exists a unique solution. Conversely, given a tangent vector ${t}^{\mu }$ at a point p of a manifold M there exists a unique geodesic which goes through p and is tangent to ${t}^{\mu }$.

1.4.3. The Riemann curvature tensor

Definition. The parallel transport of a vector from point p to point q on the manifold M is actually path-dependent. This path dependence is directly measured by the so-called Riemann curvature tensor. The Riemann curvature tensor can be defined in terms of the failure of successive operations of differentiation to commute. Let us start with an arbitrary tangent dual vector ${\omega }_{a}$ and an arbitrary function f. We want to calculate $({{\rm{\nabla }}}_{a}{{\rm{\nabla }}}_{b}-{{\rm{\nabla }}}_{b}{{\rm{\nabla }}}_{a}){\omega }_{c}$. First we have

Equation (1.65)

Similarly

Equation (1.66)

Thus

Equation (1.67)

We can follow the same set of arguments which led from equations ()–() to conclude that the tensor $({{\rm{\nabla }}}_{a}{{\rm{\nabla }}}_{b}-{{\rm{\nabla }}}_{b}{{\rm{\nabla }}}_{a}){\omega }_{c}$ depends only on the value of ${\omega }_{c}$ at the point p. In other words ${\text{}}{{\rm{\nabla }}}_{a}{{\rm{\nabla }}}_{b}-{\text{}}{{\rm{\nabla }}}_{b}{{\rm{\nabla }}}_{a}$ is a linear map which takes tangent dual vectors into tensors of type (0,3). Equivalently we can say that the action of ${\text{}}{{\rm{\nabla }}}_{a}{{\rm{\nabla }}}_{b}-{\text{}}{{\rm{\nabla }}}_{b}{{\rm{\nabla }}}_{a}$ on tangent dual vectors is equivalent to the action of a tensor of type (1,3). Thus we can write

Equation (1.68)

The tensor ${R}_{{abc}}^{d}$ is precisely the Riemann curvature tensor. We compute explicitly

Equation (1.69)

Thus

Equation (1.70)

We obtain then the components

Equation (1.71)

The action on tangent vectors can be found as follows. Let ta be an arbitrary tangent vector. The scalar product ${t}^{a}{\omega }_{a}$ is a function on the manifold and thus

Equation (1.72)

This leads immediately to

Equation (1.73)

Generalization of this result and the previous one to higher order tensors is given by the following equation

Equation (1.74)

Properties. We state without proof the following properties of the curvature tensor:

  • •  
    Anti-symmetry in the first two indices:
    Equation (1.75)
  • •  
    Anti-symmetrization of the first three indices yields 0:
    Equation (1.76)
  • •  
    Anti-symmetry in the last two indices:
    Equation (1.77)
  • •  
    Symmetry if the pair consisting of the first two indices is exchanged with the pair consisting of the last two indices:
    Equation (1.78)
  • •  
    Bianchi identity:
    Equation (1.79)
  • •  
    The so-called Ricci tensor Rac , which is the trace part of the Riemann curvature tensor, is symmetric, namely
    Equation (1.80)
  • •  
    The Einstein tensor can be constructed as follows. By contracting the Bianchi identity and using ${{\rm{\nabla }}}_{a}{g}_{{bc}}=0$ we obtain
    Equation (1.81)
    By contracting now the two indices b and d we obtain
    Equation (1.82)
    This can be put in the form
    Equation (1.83)
    The tensor Gab is called Einstein tensor and is given by
    Equation (1.84)
    The so-called scalar curvature R is defined by
    Equation (1.85)

1.5. The stress-energy-momentum tensor

1.5.1. The stress-energy-momentum tensor

We will mostly be interested in continuous matter distributions which are extended macroscopic systems composed of a large number of individual particles. We will think of such systems as fluids. The energy, momentum, and pressure of fluids are encoded in the stress–energy–momentum tensor ${T}^{\mu \nu }$, which is a symmetric tensor of type $(2,0)$. The component ${T}^{\mu \nu }$ of the stress–energy–momentum tensor is defined as the flux of the component ${p}^{\mu }$ of the 4-vector energy–momentum across a surface of constant ${x}^{\nu }$.

Let us consider an infinitesimal element of the fluid in its rest frame. The spatial diagonal component Tii is the flux of the momentum pi across a surface of constant xi , i.e. it is the amount of momentum pi per unit time per unit area traversing the surface of constant xi . Thus Tii is the normal stress, which we also call pressure when it is independent of direction. We write ${T}^{{ii}}={P}_{i}$. The spatial off-diagonal component Tij is the flux of the momentum pi across a surface of constant xj , i.e. it is the amount of momentum pi per unit time per unit area traversing the surface of constant xj , which means that Tij is the shear stress.

The component T00 is the flux of the energy p0 through the surface of constant x0, namely it is the amount of energy per unit volume at a fixed instant of time. Thus T00 is the energy density, i.e. ${T}^{00}=\rho {c}^{2}$, where ρ is the rest-mass density. Similarly, Ti0 is the flux of the momentum pi through the surface of constant x0, i.e. it is the i momentum density times c. The T0i is the energy flux through the surface of constant xi divided by c. They are equal by virtue of the symmetry of the stress–energy–momentum tensor, ${T}^{0i}={T}^{i0}$.

1.5.2. Perfect fluid

We begin with the case of 'dust', which is a collection of a large number of particles in spacetime at rest with respect to each other. The particles are assumed to have the same rest mass m. The pressure of the dust is obviously 0 in any direction since there is no motion of the particles, i.e. the dust is a pressureless fluid. The 4-vector velocity of the dust is the constant 4-vector velocity ${U}^{\mu }$ of the individual particles. Let n be the number density of the particles, i.e. the number of particles per unit volume as measured in the rest frame. Clearly ${N}^{i}={{nU}}^{i}=n(\gamma {u}_{i})$ is the flux of the particles, i.e. the number of particles per unit area per unit time in the xi -direction. The 4-vector number-flux of the dust is defined by

Equation (1.86)

The rest-mass density of the dust in the rest frame is clearly given by $\rho ={nm}$. This rest-mass density times c2 is the $\mu =0$, $\nu =0$ component of the stress–energy–momentum tensor ${T}^{\mu \nu }$ in the rest frame. We remark that $\rho {c}^{2}={{nmc}}^{2}$ is also the $\mu =0$, $\nu =0$ component of the tensor ${N}^{\mu }{p}^{\nu }$, where ${N}^{\mu }$ is the 4-vector number-flux and ${p}^{\mu }$ is the 4-vector energy–momentum of the dust. We define therefore the stress–energy–momentum tensor of the dust by

Equation (1.87)

The next fluid of paramount importance is the so-called perfect fluid. This is a fluid determined completely by its energy density ρ and its isotropic pressure P in the rest frame. Hence ${T}^{00}=\rho {c}^{2}$ and ${T}^{{ii}}=P$. The shear stresses Tij ($i\ne j$) are absent for a perfect fluid in its rest frame. It is not difficult to convince ourselves that stress–energy–momentum tensor ${T}^{\mu \nu }$ is given in this case in the rest frame by

Equation (1.88)

This is a covariant equation and thus it must also hold, by the principle of minimal coupling (see below), in any other global inertial reference frame. We give the following examples:

  • •  
    Dust: P = 0.
  • •  
    Gas of photons: $P=\rho {c}^{2}/3$.
  • •  
    Vacuum energy: $P=-\rho {c}^{2}\iff {T}^{{ab}}=-\rho {c}^{2}{\eta }^{{ab}}$.

1.5.3. Conservation law

The stress-energy-momentum tensor ${T}^{\mu \nu }$ is symmetric, i.e. ${T}^{\mu \nu }={T}^{\nu \mu }$. It must also be conserved:

Equation (1.89)

This should be thought of as the equation of motion of the perfect fluid. Explicitly this equation reads

Equation (1.90)

We project this equation along the 4-vector velocity by contracting it with ${U}_{\nu }$. We obtain (using ${U}_{\nu }{\partial }_{\mu }{U}^{\nu }=0$)

Equation (1.91)

We project the above equation along a direction orthogonal to the 4-vector velocity by contracting it with ${P}_{\nu }^{\mu }$ given by

Equation (1.92)

Indeed, we can check that ${P}_{\nu }^{\mu }{P}_{\lambda }^{\nu }={P}_{\lambda }^{\mu }$ and ${P}_{\nu }^{\mu }{U}^{\nu }=0$. By contracting equation (1.90) with ${P}_{\nu }^{\lambda }$ we obtain

Equation (1.93)

We consider now the non-relativistic limit defined by

Equation (1.94)

The parallel equation (1.91) becomes the continuity equation given by

Equation (1.95)

The orthogonal equation (1.93) becomes Euler's equation of fluid mechanics given by

Equation (1.96)

1.5.4. Minimal coupling

The laws of physics in general relativity can be derived from the laws of physics in special relativity by means of the so-called principle of minimal coupling. This consists in writing the laws of physics in special relativity in tensor form and then replacing the flat metric ${\eta }_{\mu \nu }$ with the curved metric ${g}_{\mu \nu }$ and the derivative operator ${\partial }_{\mu }$ with the covariant derivative operator ${{\rm{\nabla }}}_{\mu }$. This recipe works in most cases.

For example take the geodesic equation describing a free particle in special relativity given by ${U}^{\mu }{\partial }_{\mu }{U}^{\nu }=0$. Geodesic motion in general relativity is given by ${U}^{\mu }{{\rm{\nabla }}}_{\mu }{U}^{\nu }=0$. These are the geodesics of the curved metric ${g}_{\mu \nu }$ and they describe freely falling bodies in the corresponding gravitational field.

The second example is the equation of motion of a perfect fluid in special relativity which is given by the conservation law ${\partial }^{\nu }{T}_{\nu \lambda }=0$. In general relativity this conservation law becomes

Equation (1.97)

Also, by applying the principle of minimal coupling, the stress–energy–momentum tensor ${T}_{\mu \nu }$ of a perfect fluid in general relativity is given by equation (1.88) with the replacement $\eta \to g$, i.e.

Equation (1.98)

1.6. Einstein's equation

Although local gravitational forces cannot be measured by the principle of equivalence, i.e. since the spacetime manifold is locally flat, relative gravitational forces—the so-called tidal gravitational forces—can still be measured by observing the relative acceleration of nearby geodesics. This effect is described by the geodesic deviation equation.

1.6.1. Tidal gravitational forces

Let us first start by describing tidal gravitational forces in Newtonian physics. The force of gravity exerted by an object of mass M on a particle of mass m a distance r away is $\vec{F}=-\hat{r}{GMm}/{r}^{2}$, where $\hat{r}$ is the unit vector pointing from M to m and r is the distance between the center of M and m. The corresponding acceleration is $\vec{a}=-\hat{r}{GM}/{r}^{2}=-\vec{{\rm{\nabla }}}{\rm{\Phi }}$, ${\rm{\Phi }}=-{GM}/r$. We assume now that the mass m is spherical of radius ${\rm{\Delta }}r$. The distance between the center of M and the center of m is r. The force of gravity exerted by the mass M on a particle of mass dm a distance $r\pm {\rm{\Delta }}r$ away on the line joining the centers of M and m is given by $\vec{F}=-\hat{r}{GMdm}/{(r\pm {\rm{\Delta }}r)}^{2}$. The corresponding acceleration is

Equation (1.99)

The first term is precisely the acceleration experienced at the center of the body m due to M. This term does not affect the observed acceleration of particles on the surface of m. In other words, since m and everything on its surface are in a state of free fall with respect to M, the acceleration of dm with respect to m is precisely the so-called tidal acceleration, and is given by the second term in the above expansion, i.e.

Equation (1.100)

1.6.2. Geodesic deviation equation

In a flat Euclidean geometry two parallel lines always remain parallel. This is not true in a curved manifold. To see this more carefully we consider a one-parameter family of geodesics ${\gamma }_{s}(t)$ which are initially parallel and see what happens to them as we move along these geodesics when we increase the parameter t. The map $(t,s)\to {\gamma }_{s}(t)$ is smooth, one-to-one, and its inverse is smooth, which means in particular that the geodesics do not cross. These geodesics will then generate a two-dimensional surface on the manifold M. The parameters t and s can therefore be chosen to be the coordinates on this surface. This surface is given by the entirety of the points ${x}^{\mu }(s,t)\in M$. The tangent vector to the geodesics is defined by

Equation (1.101)

This satisfies therefore the equation ${T}^{\mu }{{\rm{\nabla }}}_{\mu }{T}^{\nu }=0$. The so-called deviation vector is defined by

Equation (1.102)

The product ${S}^{\mu }{ds}$ is the displacement vector between two infinitesimally nearby geodesics. The vectors ${T}^{\mu }$ and ${S}^{\mu }$ commute because they are basis vectors. Hence we must have ${[T,S]}^{\mu }={T}^{\nu }{{\rm{\nabla }}}_{\nu }{S}^{\mu }-{S}^{\nu }{{\rm{\nabla }}}_{\nu }{T}^{\mu }=0$ or equivalently

Equation (1.103)

This can be checked directly by using the definition of the covariant derivative and the way it acts on tangent vectors and equations (1.101) and (1.102).

The quantity ${V}^{\mu }={T}^{\nu }{{\rm{\nabla }}}_{\nu }{S}^{\mu }$ expresses the rate of change of the deviation vector along a geodesic. We will call ${V}^{\mu }$ the relative velocity of infinitesimally nearby geodesics. Similarly the relative acceleration of infinitesimally nearby geodesics is defined by ${A}^{\mu }={T}^{\nu }{{\rm{\nabla }}}_{\nu }{V}^{\mu }$. We compute

Equation (1.104)

This is the geodesic deviation equation. The relative acceleration of infinitesimally nearby geodesics is 0 if and only if ${R}_{{\nu }{\lambda }{\sigma }}^{\mu }=0$. Geodesics will accelerate towards, or away from, each other if and only if ${R}_{{\nu }{\lambda }{\sigma }}^{\mu }\ne 0$. Thus initially parallel geodesics with ${V}^{\mu }=0$ will fail generically to remain parallel.

1.6.3. Einsetin's equation

We will assume that, in general relativity, the tidal acceleration of two nearby particles is precisely the relative acceleration of infinitesimally nearby geodesics given by equation (1.104), i.e.

Equation (1.105)

This suggests, by comparing with equation (1.100), we make the following correspondence

Equation (1.106)

Thus

Equation (1.107)

By using the Poisson's equation (1.1) we obtain then the correspondence

Equation (1.108)

From the other hand, the stress–energy–momentum tensor ${T}^{\mu \nu }$ provides the correspondence

Equation (1.109)

We expect therefore an equation of the form

Equation (1.110)

This is the original equation proposed by Einstein. However, it has the following problem. From the fact that ${{\rm{\nabla }}}^{\nu }{G}_{\nu \sigma }=0$, we immediately obtain ${{\rm{\nabla }}}^{\nu }{R}_{\nu \sigma }={{\rm{\nabla }}}_{\sigma }R/2$, and as a consequence ${{\rm{\nabla }}}^{\nu }{T}_{\nu \sigma }={c}^{4}{{\rm{\nabla }}}_{\sigma }R/8\pi G$. This result is in direct conflict with the requirement of the conservation of the stress–energy–momentum tensor given by ${{\rm{\nabla }}}^{\nu }{T}_{\nu \sigma }=0$. An immediate solution is to consider instead the equation

Equation (1.111)

The conservation of the stress–energy–momentum tensor is now guaranteed. Furthermore, this equation is still in accord with the correspondence ${R}_{\nu \sigma }{U}^{\nu }{U}^{\sigma }\leftrightarrow 8\pi G\rho $. Indeed, by using the result $R=-4\pi {GT}/{c}^{4}$ we can rewrite the above equation as

Equation (1.112)

We compute ${R}_{\mu \nu }{U}^{\mu }{U}^{\nu }=(8\pi G/{c}^{4})({T}_{\mu \nu }{U}^{\mu }{U}^{\nu }+{c}^{2}T/2)$. By keeping only the $\mu =0$, $\nu =0$ component of ${T}_{\mu \nu }$ and neglecting the other components, the right-hand side is exactly $4\pi G\rho $ as it should be.

1.6.4. Newtonian limit

The Newtonian limit of general relativity is defined by the following three requirements:

  • 1)  
    The particles are moving slowly compared with the speed of light.
  • 2)  
    The gravitational field is weak so that the curved metric can be expanded about the flat metric.
  • 3)  
    The gravitational field is static.

Geodesic equation. We begin with the geodesic equation, with the proper time τ as the parameter of the geodesic, which is

Equation (1.113)

The assumption that particles are moving slowly compared to the speed of light means that

Equation (1.114)

The geodesic equation becomes

Equation (1.115)

We recall the Christoffel symbols

Equation (1.116)

Since the gravitational field is static we have

Equation (1.117)

The second assumption that the gravitational field is weak allows us to decompose the metric as

Equation (1.118)

Thus

Equation (1.119)

The geodesic equation becomes

Equation (1.120)

In terms of components this reads

Equation (1.121)

Equation (1.122)

The first equation says that ${dt}/d\tau $ is a constant. The second equation reduces to

Equation (1.123)

Einstein's equations. Now we turn to the Newtonian limit of Einstein's equation ${R}_{\nu \sigma }=8\pi G({T}_{\nu \sigma }-\frac{1}{2}{g}_{\nu \sigma }T)/{c}^{4}$ with the stress–energy–momentum tensor ${T}_{\mu \nu }$ of a perfect fluid as a source. The perfect fluid is describing the Earth or the Sun. The stress–energy–momentum tensor is given by ${T}_{\mu \nu }=(\rho +P/{c}^{2}){U}_{\mu }{U}_{\nu }+{{Pg}}_{\mu \nu }$. In the Newtonian limit this can be approximated by the stress–energy–momentum tensor of dust given by ${T}_{\mu \nu }=\rho {U}_{\mu }{U}_{\nu }$ since in this limit pressure can be neglected as it comes from motion which is assumed to be slow. In the rest frame of the perfect fluid we have ${U}^{\mu }=({U}^{0},0,0,0)$ and since ${g}_{\mu \nu }{U}^{\mu }{U}^{\nu }=-{c}^{2}$ we obtain ${U}^{0}=c(1+{h}_{00}/2)$ and ${U}_{0}=c(-1+{h}_{00}/2)$, and as a consequence

Equation (1.124)

The inverse metric is obviously given by ${g}^{00}=-1-{h}_{00}$ since ${g}^{\mu \nu }{g}_{\nu \rho }={\delta }_{\rho }^{\mu }$. Hence

Equation (1.125)

The $\mu =0$, $\nu =0$ component of Einstein's equation is therefore

Equation (1.126)

We recall the Riemann curvature tensor and the Ricci tensor

Equation (1.127)

Equation (1.128)

Thus (using in particular ${R}_{000}^{0}=0$)

Equation (1.129)

The Christoffel symbols are linear in the metric perturbation and thus one can neglect the third and fourth terms in the above equation. We then obtain

Equation (1.130)

Einstein's equation reduces therefore to Newton's equation, i.e.

Equation (1.131)

1.7. Killing vectors and maximally symmetric spaces

A spacetime which is spatially homogeneous and spatially isotropic is a spacetime in which the space is maximally symmetric. A maximally symmetric space is a space with the maximum number of isometries, i.e. the maximum number of symmetries of the metric. These isometries are generated by the so-called Killing vectors.

As an example, if ${\partial }_{\sigma }{g}_{\mu \nu }=0$, for some fixed value of σ, then the translation ${x}^{\sigma }\to {x}^{\sigma }+{a}^{\sigma }$ is a symmetry and thus it is an isometry of the curved manifold M with metric ${g}_{\mu \nu }$. This symmetry will be naturally associated with a conserved quantity. To see this let us first recall that the geodesic equation can be rewritten in terms of the 4-vector energy–momentum ${p}^{\mu }={{mU}}^{\mu }$ as ${p}^{\mu }{{\rm{\nabla }}}_{\mu }{p}_{\nu }=0$. Explicitly

Equation (1.132)

Thus if the metric is invariant under the translation ${x}^{\sigma }\to {x}^{\sigma }+{a}^{\sigma }$ then ${\partial }_{\sigma }{g}_{\mu \nu }=0$, and as a consequence the momentum ${p}_{\sigma }$ is conserved as expected.

For obvious reasons we must rewrite the condition which expresses the symmetry under ${x}^{\sigma }\to {x}^{\sigma }+{a}^{\sigma }$ in a covariant fashion. Let us thus introduce the vector $K={\partial }_{(\sigma )}$ via its components which are given (in the basis in which ${\partial }_{\sigma }{g}_{\mu \nu }=0$) by

Equation (1.133)

Clearly then ${p}_{\sigma }={p}_{\mu }{K}^{\mu }$. Since ${\partial }_{\sigma }{g}_{\mu \nu }=0$ we must have ${{dp}}_{\sigma }/{dt}=0$ or equivalently $d({p}_{\mu }{K}^{\mu })/{dt}=0$. This means that the directional derivative of the scalar quantity ${p}_{\mu }{K}^{\mu }$ along the geodesic is 0, i.e.

Equation (1.134)

We compute

Equation (1.135)

We obtain therefore the so-called Killing equation

Equation (1.136)

Thus for any vector K which satisfies the Killing equation ${{\rm{\nabla }}}_{\mu }{K}_{\nu }+{{\rm{\nabla }}}_{\nu }{K}_{\mu }=0$ the momentum ${p}_{\mu }{K}^{\mu }$ is conserved along the geodesic with tangent p. The vector K is called a Killing vector. The Killing vector K generates the isometry which is associated with the conservation of ${p}_{\mu }{K}^{\mu }$. The symmetry transformation under which the metric is invariant is expressed as infinitesimal motion in the direction of K.

Let us check that the vector ${K}^{\mu }={\delta }_{\sigma }^{\mu }$ satisfies the Killing equation. Immediately, we have ${K}_{\mu }={g}_{\mu \sigma }$ and

Equation (1.137)

Thus if the metric is independent of ${x}^{\sigma }$ then the vector ${K}^{\mu }={\delta }_{\sigma }^{\mu }$ will satisfy the Killing equation. Conversely if a vector satisfies the Killing equation then one can always find a basis in which the vector satisfies ${K}^{\mu }={\delta }_{\sigma }^{\mu }$. However, if we have more than one Killing vector we cannot find a single basis in which all of them satisfy ${K}^{\mu }={\delta }_{\sigma }^{\mu }$.

Some of the properties of Killing vectors are:

Equation (1.138)

Equation (1.139)

Equation (1.140)

The last identity in particular shows explicitly that the geometry does not change under a Killing vector.

The isometries of Rn with flat Euclidean metric are n independent translations and $n(n-1)/2$ independent rotations (which form the group of SO(n) rotations). Hence Rn with flat Euclidean metric has $n+n(n-1)/2=n(n+1)/2$ isometries. This is the number of Killing vectors on Rn with flat Euclidean metric, which is the maximum possible number of isometries in n dimensions. The space Rn is therefore called maximally symmetric space. In general a maximally symmetric space is any space with $n(n+1)/2$ Killing vectors (isometries). These spaces have the maximum degree of symmetry. The only Euclidean maximally symmetric spaces are planes Rn with 0 scalar curvature, spheres Sn with positive scalar curvature and hyperboloids Hn with negative scalar curvature 1 .

The curvature of a maximally symmetric space must be the same everywhere (translations) and the same in every direction (rotations). More precisely, a maximally symmetric space must be locally fully characterized by a constant scalar curvature R and furthermore must look like the same in all directions, i.e. it must be invariant under all Lorentz transformations at the point of consideration.

In the neighborhood of a point $p\in M$ we can always choose an inertial reference frame in which ${g}_{\mu \nu }={\eta }_{\mu \nu }$. This is invariant under Lorentz transformations at p. Since the space is maximally symmetric the Riemann curvature tensor ${R}_{\mu \nu \lambda \rho }$ at p must also be invariant under Lorentz transformations at p. This tensor must therefore be constructed from ${\eta }_{\mu \nu }$, the Kronecker delta ${\delta }_{\mu \nu }$ and the Levi-Civita tensor ${\varepsilon }_{\mu \nu \lambda \rho }$, which are the only tensors which are known to be invariant under Lorentz transformations. However, the curvature tensor satisfies ${R}_{\mu \nu \lambda \gamma }=-{R}_{\nu \mu \lambda \gamma }$, ${R}_{\mu \nu \lambda \gamma }=-{R}_{\mu \nu \gamma \lambda }$, ${R}_{\mu \nu \lambda \gamma }={R}_{\lambda \gamma \mu \nu }$, ${R}_{[\mu \nu \lambda ]\gamma }=0$, and ${{\rm{\nabla }}}_{[\mu }{R}_{\nu \lambda ]\gamma \rho }=0$. The only combination formed out of ${\eta }_{\mu \nu }$, ${\delta }_{\mu \nu }$, and ${\varepsilon }_{\mu \nu \lambda \rho }$ which satisfies these identities is ${R}_{\mu \nu \lambda \gamma }=\kappa ({\eta }_{\mu \lambda }{\eta }_{\nu \gamma }-{\eta }_{\mu \gamma }{\eta }_{\nu \lambda })$, with κ a constant. This tensorial relation must hold in any other coordinate system, i.e.

Equation (1.141)

We compute ${R}_{{\mu }{\nu }{\lambda }}^{\gamma }=\kappa ({g}_{\mu \lambda }{\delta }_{\nu }^{\gamma }-{\delta }_{\mu }^{\gamma }{g}_{\nu \lambda })$, ${R}_{\mu \lambda }={R}_{{\mu }{\nu }{\lambda }}^{\gamma }=\kappa (n-1){g}_{\mu \lambda }$, and hence $R=\kappa n(n-1)$. In other words the scalar curvature of a maximally symmetric space is a constant over the manifold. Thus the curvature of a maximally symmetric space must be of the form

Equation (1.142)

Conversely if the curvature tensor is given by this equation with R constant over the manifold then the space is maximally symmetric.

1.8. The Hilbert–Einstein action

Einstein's equation for general relativity reads

Equation (1.143)

The dynamical variable is obviously the metric ${g}_{\mu \nu }$. The goal is to construct an action principle from which the Einstein's equations follow as the Euler–Lagrange equations of motion for the metric. This action principle will read as

Equation (1.144)

The first problem with this way of writing is that both dnx and ${ \mathcal L }$ are tensor densities rather than tensors. We digress briefly to explain this important difference.

Let us recall the familiar Levi-Civita symbol in n dimensions defined by

Equation (1.145)

This is a symbol and not a tensor since it does not change under coordinate transformations. The determinant of a matrix M can be given by the formula

Equation (1.146)

By choosing ${M}_{\nu }^{\mu }=\partial {x}^{\mu }/\partial {y}^{{\nu }^{}}$ we obtain the transformation law

Equation (1.147)

In other words ${\tilde{\varepsilon }}_{{\mu }_{1}^{}\ldots {\mu }_{n}^{}}$ is not a tensor because of the determinant appearing in this equation. This is an example of a tensor density. Another example of a tensor density is $\mathrm{det} g$. Indeed from the tensor transformation law of the metric ${g}_{\alpha \beta }^{^{\prime} }={g}_{\mu \nu }(\partial {x}^{\mu }/\partial {y}^{\alpha })(\partial {x}^{\nu }/\partial {y}^{\beta })$ we can show in a straightforward way that

Equation (1.148)

The actual Levi-Civita tensor can then be defined by

Equation (1.149)

Next, under a coordinate transformation $x\to y$ the volume element transforms as

Equation (1.150)

In other words the volume element transforms as a tensor density and not as a tensor. We verify this important point in our language as follows. We write

Equation (1.151)

Recall that a differential p-form is a $(0,p)$ tensor which is completely antisymmetric. For example scalars are 0-forms and dual cotangent vectors are 1-forms The Levi-Civita tensor ${\varepsilon }_{{\mu }_{1}\ldots {\mu }_{n}}$ is a 4-form. The differentials ${{dx}}^{\mu }$ appearing in the second line of equation (1.151) are 1-forms and hence under a coordinate transformation $x\to y$ we have ${{dx}}^{\mu }\to {{dy}}^{\mu }={{dx}}^{\nu }\partial {y}^{\mu }/\partial {x}^{\nu }$. By using this transformation law we can immediately show that dxn transforms to dny exactly as in equation (1.150).

It is not difficult to see now that an invariant volume element can be given by the n-form defined by the equation

Equation (1.152)

We can show that

Equation (1.153)

In other words the invariant volume element is precisely the Levi-Civita tensor. In the case of Lorentzian signature we replace $\mathrm{det}\; g$ with $-\mathrm{det}\; g$.

We go back now to equation (1.144) and rewrite it as

Equation (1.154)

Clearly ${ \mathcal L }=\sqrt{-\mathrm{det}\; g}\hat{{ \mathcal L }}$. Since the invariant volume element ${d}^{n}x\sqrt{-\mathrm{det}\; g}$ is a scalar the function $\hat{{ \mathcal L }}$ must also be a scalar and as such can be identified with the Lagrangian density.

We use the result that the only independent scalar quantity which is constructed from the metric and which is at most second order in its derivatives is the Ricci scalar R. In other words the simplest choice for the Lagrangian density $\hat{{ \mathcal L }}$ is

Equation (1.155)

The corresponding action is called the Hilbert–Einstein action. We compute

Equation (1.156)

We have

Equation (1.157)

In the second line of the above equation we have used the fact that $\delta {{\rm{\Gamma }}}^{\rho }{}_{\mu \nu }$ is a tensor since it is the difference of two connections. Thus

Equation (1.158)

We compute also (with $\delta {g}_{\mu \nu }=-{g}_{\mu \alpha }{g}_{\nu \beta }\delta {g}^{\alpha \beta }$)

Equation (1.159)

Thus

Equation (1.160)

By Stokes's theorem this integral is equal to the integral over the boundary of spacetime of the expression ${g}_{\mu \nu }{{\rm{\nabla }}}^{\rho }\delta {g}^{\mu \nu }-{{\rm{\nabla }}}_{\mu }\delta {g}^{\mu \rho }$, which is 0 if we assume that the metric and its first derivatives are held fixed on the boundary. The variation of the action reduces to

Equation (1.161)

Next we use the result

Equation (1.162)

Hence

Equation (1.163)

This will obviously lead to Einstein's equations in a vacuum which is partially our goal. We also want to include the effect of matter, which requires considering the more general actions of the form

Equation (1.164)

Equation (1.165)

The variation of the action becomes

Equation (1.166)

In other words

Equation (1.167)

Einstein's equations are therefore given by

Equation (1.168)

The stress–energy–momentum tensor must therefore be defined by the equation

Equation (1.169)

As a first example we consider the action of a scalar field in curved spacetime given by

Equation (1.170)

The corresponding stress–energy–momentum tensor is calculated to be given by

Equation (1.171)

As a second example we consider the action of the electromagnetic field in curved spacetime given by

Equation (1.172)

In this case the stress–energy–momentum tensor is calculated to be given by

Equation (1.173)

1.9. Exercises

Exercises

Exercises Hit 'Start interaction' to show exercises.

Start interaction
Standard image

References

Export references: BibTeX RIS

Footnotes

  • 1  

    The corresponding maximally symmetric Lorentzian spaces are Minkowski spaces Mn (R = 0), de Sitter spaces dSn ($R\gt 0$), and anti-de Sitter spaces AdSn ($R\lt 0$).