6.1 An Exact Solution

The special theory of relativity assumes the existence of a unique class of global coordinate systems - called inertial coordinates - with respect to which the speed of light in vacuum is everywhere equal to the constant c. It was natural, then, to express physical laws in terms of this preferred class of coordinate systems, characterized by the global invariance of the speed of light. In addition, the special theory also strongly implied the fundamental equivalence of mass and energy, according to which light (and every other form of energy) must be regarded as possessing inertia. However, it soon became clear that the global invariance of light speed together with the idea that energy has inertia (as expressed in the famous relation E2 = m2 + |p|2) were incompatible with one of the most firmly established empirical results of physics, namely, the exact proportionality of inertial and gravitational mass, which Einstein elevated to the status of a Principle. This incompatibility led Einstein, as early as 1907, to the belief that the global invariance of light speed, in the sense of the special theory, could not be maintained. Indeed, he concluded that we cannot assume, as do both Newtonian theory and special relativity, the existence of any global inertial systems of coordinates (although we can carry over the existence of a local system of inertial coordinates in a vanishingly small region of spacetime around any event).

Since no preferred class of global coordinate systems is assumed, the general theory essentially places all (smoothly related) systems of coordinates on an equal footing, and expresses physical laws in a way that is applicable to any of these systems. As a result, the laws of physics will hold good even with respect to coordinate systems in which the speed of light takes on values other than c. For example, the laws of general relativity are applicable to a system of coordinates that is fixed rigidly to the rotating Earth. According to these coordinates the distant galaxies are "circumnavigating" nearly the entire universe in just 24 hours, so their speed is obviously far greater than the constant c. The huge implied velocities of the celestial spheres was always problematical for the ancient conception of an immovable Earth, but it is beautifully accommodated within general relativity by the effect which the implied centrifugal acceleration field - whose strength increases in direct proportion to the distance from the Earth - has on the values of the metric components guv for this rotating system of coordinates at those locations. It's true that, when expressed in this rotating system of coordinates, those stars are moving with dx/dt values that far exceed the usual numerical value of c, but they are not moving faster than light, because the speed of light at those locations, expressed in terms of those coordinates, is correspondingly greater.

In general, the velocity of light can always be inferred from the components of the metric tensor, and typically looks something like . To understand why this is so, recall that in special relativity we have

and the trajectory of a light ray follows a null path, i.e., a path with dt = 0. Thus, dividing by (dt)2, we see that the path of light through spacetime satisfies the equation

and so the velocity of light is unambiguous in the context of special relativity, which is restricted to inertial coordinate systems with respect to which equation (1) is invariant. However, in the general theory we are no longer guaranteed the existence of a global coordinate system of the simple form (1). It is true that over a sufficiently small spatial and temporal region surrounding any given point in spacetime there exists a coordinate system of that simple Minkowskian form, but in the presence of a non-vanishing gravitational field ("curvature") equation (1) applies only with respect to "free-falling" reference frames, which are necessarily transient and don't extend globally.

So, for example, instead of writing the metric in the xt plane as (dt)2 = (dt)2 - (dx)2 , we must consider the more general form

As always, the path of a light ray is null, so we have dt = 0, and the differentials dx and dt must satisfy the equation

Solving this gives

If we diagonalize our metric we get gxt = 0, in which case the "velocity" of a null path in the xt plane with respect to this coordinate system is simply dx/dt = . This quantity can (and does) take on any value, depending on our choice of coordinate systems.

Around 1911 Einstein proposed to incorporate gravitation into a modified version of special relativity by allowing the speed of light to vary as a scalar from place to place as a function of the gravitational potential. This "scalar c field" is remarkably similar to a simple refractive medium, in which the speed of light varies as a function of the density. Fermat's principle of least time can then be applied to define the paths of light rays as geodesics in the spacetime manifold (as discussed in Section 8.4). Specifically, Einstein wrote in 1911 that the speed of light at a place with the gravitational potential j would be c0 (1 + j/c02), where c0 is the nominal speed of light in the absence of gravity. In geometrical units we define c0 = 1, so Einstein's 1911 formula can be written simply as c = 1 + j. However, this formula for the speed of light (not to mention this whole approach to gravity) turned out to be incorrect, as Einstein realized during the years leading up to 1915 and the completion of the general theory. In fact, the general theory of relativity doesn't give any equation for the speed of light at a particular location, because the effect of gravity cannot be represented by a simple scalar field of c values. Instead, the "speed of light" at a each point depends on the direction of the light ray through that point, as well as on the choice of coordinate systems, so we can't generally talk about the value of c at a given point in a non-vanishing gravitational field. However, if we consider just radial light rays near a spherically symmetrical (and non- rotating) mass, and if we agree to use a specific set of coordinates, namely those in which the metric coefficients are independent of t, then we can read a formula analogous to Einstein's 1911 formula directly from the Schwarzschild metric. But how does the Schwarzschild metric follow from the field equations of general relativity?

To deduce the implications of the field equations for observable phenomena Einstein originally made use of approximate methods, since no exact solutions were known. These approximate methods were adequate to demonstrate that the field equations lead in the first approximation to Newton's laws, and in the second approximation to a natural explanation for the anomalous precession of Mercury (see Section 6.2). However, these results can now be directly computed from the exact solution for a spherically symmetric field, found by Karl Schwarzschild in 1916. As Schwarzschild wrote, it's always pleasant to find exact solutions, and the simple spherically symmetrical line element "let's Mr. Einstein's result shine with increased clarity". To this day, most of the empirically observable predictions of general relativity are consequences of this simple solution.

We will discuss Schwarzschild's original derivation in Section 8.7, but for our present purposes we will take a slightly different approach. Recall from Section 5.5 that the most general form of the metrical spacetime line element for a spherically symmetrical static field (although it is not strictly necessary to assume the field is static) can be written in polar coordinates as

where gqq = -r2, gff = -r2 sin(q)2, and gtt and grr are functions of r and the gravitating mass m. We expect that if m = 0, and/or as r increases to infinity, we will have gtt = 1 and grr = -1 in order to give the flat Minkowski metric in the absence of gravity. We've seen that in this highly symmetrical context there is a very natural way to derive the metric coefficients gtt and grr simply from the requirement to satisfy Kepler's third law and the principle of symmetry between space and time. However, we now wish to know what values for these metric coefficients are implied by Einstein's field equations.

In any region that is free of (non-gravitational) mass-energy the vacuum field equations must apply, which means the Ricci tensor

must vanish, i.e., all the components are zero. Since our metric is in diagonal form, it's easy to see that the Christoffel symbols for any three distinct indices a,b,c reduce to

with no summations implied. In two of the non-vanishing cases the Christoffel symbols are of the form qa/(2q), where q is a particular metric component and subscripts denote partial differentiation with respect to xa. By an elementary identity these can also be written as . Hence if we define the new variable we can write the Christoffel symbol in the form Qa with q = e2Q. Accordingly if we define the variables (functions of r)

then we have

and the non-vanishing Christoffel symbols (as given in Section 5.5) can be written as

We can now write down the components of the Ricci tensor, each of which must vanish in order for the field equations to be satisfied. Writing them out explicitly and expanding all the implied summations for our line element, we find that all the non-diagonal components are identically zero (which we might have expected from symmetry arguments), so the only components of interest in our case are the diagonal elements

Inserting the expressions for the Christoffel symbols gives the equations for the four diagonal components of the Ricci tensor as functions of u and v:

The necessary and sufficient condition for the field equations to be satisfied by a line element of the form (2) is that these four quantities each vanish. Combining the expressions for Rtt and Rrr we immediately have ur = -vr , which implies u = -v + k for some arbitrary constant k. Making these substitutions into the equation for Rqq and setting the constant of integration to k = pi/2 gives the condition

Remembering that e2u = gtt, and that the derivative of e2u is 2ur e2u, this condition expresses the requirement

The left side is just the chain rule for the derivative of the product r gtt, and since this derivative equals 1 we immediately have rgtt = r + a for some constant a . Also, since grr = e2v where v = -u + pi/2, it follows that grr = -1/gtt, and so we have the results

To match the Newtonian limit we set a = -2m where m is classically identified with the mass of the gravitating body. These metric coefficients were derived by combining the expressions for Rtt and Rrr, but it's easy to verify that they also satisfy each of those equations separately, so this is indeed the unique spherically symmetrical static solution of Einstein's field equations.

Now that we have derived the Schwarzschild metric, we can easily correct the "speed of light" formula that Einstein gave in 1911. A ray of light always travels along a null trajectory, i.e., with dt = 0, and for a radial ray we have dq and df both equal to zero, so the equation for the light ray trajectory through spacetime, in Schwarzschild coordinates (which are the only spherically symmetrical ones in which the metric is independent of t) is simply

from which we get

where the sign just indicates that the light can be going radially inward or outward. (Note that we're using geometric units, so c = 1.) In the Newtonian limit the classical gravitational potential at a distance r from mass m is j = - m/r, so if we let cr = dr/dt denote the radial speed of light in Schwarzschild coordinates, we have

cr = 1 + 2j

which corresponds to Einstein's 1911 equation, except that we have a factor of 2 instead of 1 on the potential term. Thus, as j becomes increasingly negative (i.e., as the magnitude of the potential increases), the radial "speed of light" cr defined in terms of the Schwarzschild parameters t and r is reduced to less than the nominal value of c.

On the other hand, if we define the tangential speed of light at a distance r from a gravitating mass center in the equatorial plane (q = p/2) in terms of the Schwarzschild coordinates as ct = r(df/dt), then the metric divided by (dt)2 immediately gives

Thus, we again find that the "velocity of light" is reduced a region with a strong gravitational field, but this speed is the square root of the radial speed at the same point, and to the first order in m/r this is the same as Einstein's 1911 formula, although it is understood now to signify just the tangential speed. This illustrates the fact that the general theory doesn't lead to a simple scalar field of c values. The effects of gravitation can only be accurately represented by a tensor field.

One of the observable implications of general relativity (as well as any other theory that respects the equivalence principle) is that the rate of proper time at a fixed radial position in a gravitational field relative to the coordinate time (which corresponds to proper time sufficiently far from the gravitating mass) is given by

It follows that the characteristic frequency n 1 of light emitted by some known physical process at a radial location r1 will represent a different frequency n 1 with respect to the proper time at some other radial location r2 according to the formula

From the Schwarzschild metric we have gtt(rj) = 1+2jj where jj = -m/rj is the gravitational potential at rj, so

Neglecting the higher-order terms and rearranging, this can also be written as

Observations of the light emitted from the surface of the Sun, and from other stars, is consistent with this predicted amount of gravitational redshift (up to first order), although measurements of this slight effect are difficult. A terrestrial experiment performed by Rebka and Pound in 1960 exploited the Mossbauer effect to precisely determine the redshift between the top and bottom of a tower. The results were in good agreement with the above formula, and subsequent experiments of the same kind have improved the accuracy to within about 1 percent. (Note that if r1 and r2 are nearly equal, as, for example, at two heights near the Earth's surface, then the leading factor of the right-most expression is essentially just the acceleration of gravity a = -m/r2, and the factor in parentheses is the difference in heights Dh, so we have Dn/n = a Dh.)

However, it's worth noting that this amount of gravitational redshift is a feature of just about any viable theory of gravity that includes the equivalence principle, so these experimental results, although useful for validating that principle, are not very robust for distinguishing between competing theories of gravity. For this we need to consider other observations, such as the paths of light near a gravitating body, and the precise orbits of planets. These phenomena are discussed in the subsequent sections.

Return to Table of Contents

Сайт управляется системой uCoz