PHAS0067: Advanced Physical Cosmology

2 The Metric

()

2.1 The cosmological principle

On the largest scales, the universe is assumed to be uniform. This idea is called the cosmological principle. There are two aspects of the cosmological principle:

  • the universe is homogeneous. There is no preferred observing position in the universe.

  • the universe is isotropic. The universe looks the same in every direction.

Exercise 2A To illustrate the difference between homogeneous and isotropic, draw a sketch of a universe that’s homogeneous but not isotropic; and another sketch of a universe that’s isotropic but not homogeneous.

There is an overwhelming amount of observational evidence that the universe is expanding. This means that early in the history of the universe, the distant galaxies were closer to us than they are today. It is convenient to describe the scaling of the coordinate grid in an expanding universe by the scale factor. In a smooth, expanding universe, the scale factor connects the coordinate distance with the physical distance.

In general, we need to separate the idea of coordinates from physical distances. In normal cartesian coordinates, these two ideas are almost the same. But more generally they are different. Think about polar coordinates, for example; the angle θ is not a distance. But it can be turned into a distance using a suitable formula.

In fact a formula turning coordinate differences into a physical distance is known as a metric. The metric is an essential tool to make quantitative predictions in an expanding universe.

2.2 Example metrics

Figure 2: The metric in 2D, (left panel) for two familiar coordinate systems and (right panel) a third, non-orthogonal example. For Cartesian coordinates the distance between points A and B is just given by the Pythagoras rule: the distance ds is related to the coordinate changes dx and dy by ds2=dx2+dy2. For polar coordinates, we can still construct a right-angled triangle and apply Pythagoras, but care is required because the length of the sides are dr and rdθ (not dθ). Thus ds2=dr2+r2dθ2. However, for many coordinate systems it will not even be possible to construct the right-angled triangle (right panel, which shows the example of the uv coordinate system given in the text) and an even more general form of metric is required.

The most familiar metric is that of Cartesian coordinates (Figure 2); in 2D,

ds2=dx2+dy2=i(dxi)2, (1)

where xi refers to the two components of the Cartesian coordinates, i.e. x1=x and x2=y. The content should be extremely familiar but the notation looks a little weird if you haven’t seen it before; in particular note that indices are written as superscripts for reasons that will become clear later (subscripts are reserved for a different type of index). The context should normally distinguish between x2 meaning “the square of x” and x2 meaning “the second component of the x vector”. Do be aware of this potential for confusion, though. To be clear, (dxi)2 is the square of the change in the ith coordinate, so (dx1)2=(dx)2 and (dx2)2=(dy)2 for Cartesian coordinates.

To switch to polar coordinates, we now imagine that x1=r=(x2+y2)1/2 and x2=θ=tan-1(y/x). Thinking back to previous courses on polar coordinates, it should be clear that calculating a physical distance ds now requires a different approach:

ds2=dr2+r2dθ2i(dxi)2. (2)

In fact, things can get more messy still. Suppose I define a new coordinate system (u,v) where u=2x and v=y-x. The right panel of Figure 2 shows how the displacements du and dv no longer form a right-angled triangle at all. How could we calculate the physical distance ds if forced to work in these coordinates?

The trick is to relate it back to what we do know, the cartesian coordinate system. We can immediately find that du=2dx and dv=(dy-dx). Inverting this relationship, we find

dx=12duanddy=12du+dv, (3)

which we can substitute into equation (1) to obtain

ds2=du2+dv2+2dudv. (4)

Note the appearance of cross-terms in dudv; this is the sign of a non-orthogonal coordinate system because it violates Pythagoras’ theorem. But don’t be fooled: so far there is no hint of “curvature”, which is an intrinsic property of spaces and not dependent on the coordinate system in use.

We can generalise rules (1), (2) and (4) using the concept of the metric tensor gij. In 2D (Fig. 2),

ds2=i,j=1,2gijdxidxj, (5)

where the metric gij is a 2×2 symmetric matrix. Note that this allows for cross-terms of the form seen in Equation (4). In Cartesian coordinates,

x1=x,x2=y,gij=(1001), (6)

while in polar coordinates,

x1=r,x2=θ,gij=(100r2). (7)

For our example uv coordinate system,

x1=u,x2=v,gij=(112121). (8)

Note that gij is defined to be symmetric in ij.

Exercise 2B Check that substituting equations (6), (7) and (8) into (5) correctly regenerates the expressions (1), (2) and (4) for the distances in these example coordinate systems. Optional: Show that the concept of a metric tensor expresses a completely general way of describing distances in any coordinate system that can be written as a non-singular function of the cartesian coordinates. (In particular, you may like to show that all terms must be quadratic in the coordinate displacements.)

In classical Newtonian mechanics, gravity is a force, and particles travel along curved trajectories because of this force. In general relativity, gravity is instead encoded in the metric; particles actually move along the closest equivalent to a straight line, but this begins to look curved because of curvature built into the space itself. To start formalising and using this concept, we need metrics in 4 dimensions (1 time + 3 space). The physically invariant spacetime distance is then defined by

ds2=μ,ν=03gμνdxμdxν, (9)

where μ,ν{0,1,2,3}. Typically 0 is chosen as the time and 13 as the space coordinates; however there is no unique choice of coordinates (not even for time) and therefore we refer to coordinate 0 as timelike and 13 as spacelike respectively.

As a final adjustment to our notation, we will adopt the Einstein summation convention which states that repeated indices in a single term are summed over. So, we can rewrite Equation (9) more compactly as:

ds2=gμνdxμdxν, (10)

2.3 Vectors and covectors

The displacement dxμ is an example of a vector. However we can generalise the concept to a generic vector Aμ. Vectors are, by definition, any object that behaves like a small displacement under transformations of the coordinate system (for example: a rotation, or a change from cartesian to polar coordinates). In particular, in a new coordinate system xμ the coordinate displacements transform via the chain rule:

dxμ=xμxνdxν, (11)

via the chain rule. (Don’t forget we are now using the Einstein summation convention so there is an implicit sum over ν from 0 to 3 – but not over μ even though it appears twice, as the appearances are in different terms). All vectors transform in the same way:

😇 Exercise 2C Any 4-vector Vμ can be thought of as an arrow connecting two very close points in spacetime – specifically, let’s imagine it connects xμ and xμ+ϵVμ for a very small ϵ1. Using a Taylor expansion show that, under the coordinate transformation xμxμ, the vector Vμ transforms according to the rule Vμ=xμxνVν, (12) Note that the inverse relation also immediately follows, Vμ=xμxνVν. (13) just by a trivial relabelling of which coordinate system we think of as the “primed” one.

In summary, one can think of vectors as little arrows living at a particular point in the spacetime. Mathematicians express this idea by saying that vectors belong to the tangent space which can be defined more formally.

A vector has a length |𝑨| given by |𝑨|2gμνAμAν. One can generalise this notion to a dot product of two vectors 𝑨𝑩gμνAμBν. These constructions are all independent of the coordinate system in which they are evaluated because, under a coordinate transformation xμxμ, the metric becomes

gμν=xαxμxβxνgαβ, (14)

which follows from the invariance of ds2 in equation (10).

😇 Exercise 2D Flesh out this argument: show that gμν must transform this way given the transformation (11). Then show that, as a result, AμBνgμν is also invariant under coordinate transformations for any vectors 𝑨 and 𝑩.

There is another way to think about the dot product: we can imagine that to each vector Aμ corresponds a covector Aμ defined by

AμgμνAν. (15)

Then the dot product 𝑨𝑩 is expressed simply as 𝑨𝑩=AμBμ. Covector indices transform under coordinate transformations as follows:

AμAμ=xνxμAν, (16)

which you can verify from the definition (15) and existing transformation laws (12) and (14).

Finally, it is sometimes useful to be able to make a vector again starting from a covector. For this purpose we use the inverse metric gμν:

Aμ=gμνAνwheregαβgβγ=δγα. (17)

Here δγα is the Kronecker delta:

δγα={1(α=γ)0(αγ), (18)

i.e. the identity matrix in component form. All this does is to define gμν as the matrix inverse of gμν, so that going from a vector to a covector and then back again gives you the vector you started with.

2.4 Special relativity metric

As we will now see, the special nature of spacetime (as opposed to just ordinary space extended to four dimensions) is that space and time coordinate carry opposite signs in the metric. In this course we will use the following convention for the signature of the metric: (+,-,-,-). So, the physical distance associated with a spacelike coordinate displacement is -ds2. The time interval associated with a timelike coordinate displacement is ds2.

Beware: while this convention is commonly used by particle physicists, the convention used in relativity and cosmology is sometimes (-,+,+,+). This is one of a number of conventions in cosmology that are not universally agreed upon and leads to frustrating confusion when comparing results from different authors.

The Minkowski metric ημν is the metric of special relativity, and it describes flat space. Its line element is given by

ds2 = ημνdxμdxν, (19)
= dt2-(dx2+dy2+dz2), (20)

and so the components of the metric are

ημν=diag(1,-1,-1,-1). (21)

Note that in this course we set the speed of light c=1; otherwise c2 would appear in the first entry (generating c2dt2 as the first term in (20)). SR applies in inertial frames (where there is no gravity). We’ll shortly see that it also applies locally in GR frames that are free-falling. Because it is only locally equivalent, we can’t say in general that there is a GR frame where

Δs2=Δt2-Δx2-Δy2-Δz2 (22)

for finite coordinate displacements Δt, Δx, Δy, Δz. In fact, if equation (22) holds regardless of the size of the coordinate changes, one may show there is no gravitational field. Conversely, when a gravitational field is present, (22) cannot apply everywhere. The relationship between the individual locally flat coordinate systems in neighbouring patches is what generates curvature and gravity.

2.5 Metrics in general relativity

Let’s consider two forms of the key principle that Einstein used to develop general relativity.

Weak Equivalence Principle (WEP): At any point in a gravitational field, in a frame moving with the free fall acceleration at that point, the laws of motion of free test particles have their usual Special Relativity (SR) form, i.e. particles move in straight lines with uniform velocity locally.

Strong Equivalence Principle (SEP): At any point in a gravitational field, in a frame moving with the free fall acceleration at that point, all the laws of physics have their usual Special Relativity (SR) form, except gravity, which disappears locally.

Instead of thinking of particles moving under a force, the WEP allows us to think of them in a frame without gravity, but moving with the free fall acceleration at that point. The SEP goes further and claims that all laws of physics (e.g. those governing electromagnetic fields) also behave in this way. Clearly, the WEP follows from the SEP. Despite its name, the WEP is in itself a very powerful idea which has two important consequences. First, it explains the equivalence of gravitational mass and inertial mass. Second, it says that the motion of a test body in a gravitational field only depends on its position and instantaneous velocity in spacetime.

The equivalence principle works only locally (technically, in an infinitessimally small patch) where the gravitational field can be taken to be uniform. If the gravitational field varies across a patch, tidal forces appear which cannot be ignored. This is just like curvature – on a small enough scale, any surface looks flat. But as one takes a broader view, variations in the surface become significant.

True gravity as opposed to just an accelerating frame manifests itself via the separation or coming together of test particles initially on parallel tracks. This is equivalent to what we said above: there is no global Lorentz frame in the presence of a non-uniform gravitational field.

2.6 Metric of a spatially-flat expanding universe

Figure 3: If the comoving distance today at time t0 is x0, the physical distance between the two points at some earlier time t<t0 was a(t)x0.

What is the metric of an expanding universe?

If the distance today is x0, the physical distance between two points at some earlier time t was a(t)x0 (see Figure above). In a flat (as opposed to open or closed) universe, the spatial part of the metric must look like a normal Euclidean metric, except that the distance must be multiplied by the scale factor a(t). Thus, the metric of a flat, expanding universe is the Friedmann-Robertson-Walker metric:

gμν=diag(1,-a2(t),-a2(t),-a2(t)). (23)

or, equivalently,

ds2=dt2-a2(t)(dx2+dy2+dz2) (24)

For once, it’s worth memorising these. You probably will do so naturally by the end of the course if you work through the exercises.

When inhomogeneities are introduced, the metric will become more complicated; the perturbed part of the metric will be determined by the irregularities in the matter and the radiation.

2.7 Summary

There is some important notation to understand:

  • Indices appear in superscript for vectors, e.g. Aμ. The time component is A0, while the space components are A1, A2 and A3. Obviously this can look a bit like you’re taking powers of some scalar quantity A, but normally the context should make clear what is going on.

  • Indices appear in subscript for covectors, e.g. Bμ. The time component is B0, while the space components are B1, B2 and B3.

  • The components of vectors change when you change coordinate system; so do covectors, but the transformation law is different. For examination purposes you would not be expected to remember the exact transformation laws; but if you need to refer to them, they are given by (12) and (16) respectively.

  • When you encounter a superscript and a subscript index in the same term, it means that there is implicitly a sum, i.e. AμBμμ=03AμBμ.

  • This in turn means that AμBμ is a scalar (i.e. a single number that does not depend on coordinate system).

  • Greek indices like μ, ν etc are spacetime indices which run from 0 to 3; regular indices like a, b, i, j etc are spatial indices, running from 1 to 3.

  • “Tensors” are generalisations of vectors and covectors; they can have any number of independent indices. A tensor with two indices can be thought of as a matrix that changes in a specific way if you change coordinate system. Again, you wouldn’t be expected to remember the transformation law in an exam for this course, but it is a straightforward generalisation of the rule for covectors and vectors.

With this in hand, you need to remember that the spacetime metric gμν is a tensor that:

  • provides the connection between the values of the coordinates and the physically meaningful measure of the spacetime interval ds2, following the rule ds2=gμνdxμdxν;

  • has values depending on the coordinate system in such a way that gμνAμBν is independent of coordinate system for any vectors A and B;

  • is defined to be symmetric in μν;

  • in principle, has 4 diagonal and 6 off-diagonal components which must be specified to describe a spacetime in a particular coordinate system;

  • can convert a vector Aμ into a covector Aμ via the rule Aμ=gμνAν;

  • has an inverse, gμν which obeys gαβgβγ=δγα and consequently converts covectors back to vectors, Aμ=gμνAν;

  • has a sign difference for spacelike as opposed to timelike directions – in this course we choose the minus sign to be associated with spacelike directions;

  • contains the scalefactor (which varies with time) in a homogeneous, expanding universe – see equation (23) and (24);

As the course progresses we will need to use more complicated metrics, but for now we can move onto understanding how the metric of an expanding universe affects physical, measurable quantities.