Tolman–Oppenheimer–Volkoff equation

In astrophysics, the Tolman–Oppenheimer–Volkoff (TOV) equation constrains the structure of a spherically symmetric body of isotropic material which is in static gravitational equilibrium, as modelled by general relativity. The equation[1] is

\frac{dP(r)}{dr}=-\frac{G}{r^2}\left[\rho(r)+\frac{P(r)}{c^2}\right]\left[M(r)+4\pi r^3  \frac{P(r)}{c^2}\right]\left[1-\frac{2GM(r)}{c^2r}\right]^{-1} \;

Here, r is a radial coordinate, and ρ(r0) and P(r0) are the density and pressure, respectively, of the material at r = r0.

The equation is derived by solving the Einstein equations for a general time-invariant, spherically symmetric metric. For a solution to the Tolman–Oppenheimer–Volkoff equation, this metric will take the form[1]

ds^2=e^{\nu(r)} c^2 dt^2 - (1-2GM(r)/rc^2)^{-1} dr^2 - r^2(d\theta^2 + \sin^2 \theta d\phi^2) \;

where ν(r) is determined by the constraint[1]

\frac{d\nu(r)}{dr}=- \left(\frac{2}{P(r)+\rho(r)c^2} \right) \frac{dP(r)}{dr} \;

When supplemented with an equation of state, F(ρ, P) = 0, which relates density to pressure, the Tolman–Oppenheimer–Volkoff equation completely determines the structure of a spherically symmetric body of isotropic material in equilibrium. If terms of order 1/c2 are neglected, the Tolman–Oppenheimer–Volkoff equation becomes the Newtonian hydrostatic equation, used to find the equilibrium structure of a spherically symmetric body of isotropic material when general-relativistic corrections are not important.

If the equation is used to model a bounded sphere of material in a vacuum, the zero-pressure condition P(r) = 0 and the condition exp[ν(r)] = 1 − 2GM(r)/rc2 should be imposed at the boundary. The second boundary condition is imposed so that the metric at the boundary is continuous with the unique static spherically symmetric solution to the vacuum field equations, the Schwarzschild metric:

ds^2=(1-2GM_0/rc^2) c^2 dt^2 - (1-2GM_0/rc^2)^{-1} dr^2 - r^2(d\theta^2 + \sin^2 \theta d\phi^2) \;

Total mass

M(r0) is the total mass inside radius r = r0, as measured by the gravitational field felt by a distant observer, it satisfies M(0) = 0.[1]

\frac{dM(r)}{dr}=4 \pi \rho(r) r^2 \;

Here, M0 is the total mass of the object, again, as measured by the gravitational field felt by a distant observer. If the boundary is at r = rB, continuity of the metric and the definition of M(r) require that

M_0=M(r_B)=\int_0^{r_B} 4\pi \rho(r) r^2 \; dr \;

Computing the mass by integrating the density of the object over its volume, on the other hand, will yield the larger value

M_1=\int_0^{r_B} \frac{4\pi \rho(r) r^2}{\sqrt{1-2GM(r)/rc^2}} \; dr \;

The difference between these two quantities,

\delta M=\int_0^{r_B} 4\pi \rho(r) r^2(1-(1-2GM(r)/rc^2)^{-1/2}) \; dr \;

will be the gravitational binding energy of the object divided by c2 and it is negative.

Derivation from general relativity

Derivation of the Oppenheimer-Volkoff equation (O-V)...

Non-rotating universal metric element functions (J = 0):

c^{2} d\tau^{2} = g_{00} c^2 dt^{2} + g_{11}dr^{2} + g_{22}d\theta^{2} + g_{33}d\phi^2 \;

Schwarzschild metric:

c^{2} d\tau^{2} = e^{\nu(r)} c^2 dt^{2} - e^{\lambda(r)} dr^{2} - r^2 d\theta^{2} - r^2 \sin^2 \theta d\phi^2 \;

Where Schwarzschild metric Einstein tensor metric element functions are defined as:

g_{00} = e^{\nu(r)} \; and g_{11} = - e^{\lambda(r)} \;

Stress-energy tensor Schwarzschild field hydrostatic density:

T_{00} = \rho(r) c^2 g_{00} = \rho(r) e^{\nu(r)} c^2 \;

Stress-energy tensor Schwarzschild field hydrostatic pressure:

T_{11} = P(r) g_{11} = - P(r) e^{\lambda(r)} \;

Where \rho(r) is the fluid density and P(r) is the fluid pressure.

Schwarzschild-Einstein tensor element derived from the Schwarzschild metric:

G_{11} = \frac{- r \nu'(r) + e^{\lambda(r)} - 1}{r^2} \;

Einstein field equation:

\frac{8 \pi G}{c^4} T_{11} = G_{11} \;

Integration via substitution:

- \frac{8 \pi G}{c^4} P(r) e^{\lambda(r)} = \frac{- r \nu'(r) + e^{\lambda(r)} - 1}{r^2} \;

Solve for \nu'(r):

\frac{d\nu(r)}{dr} = \frac{1}{r} \left( \frac{8 \pi G r^2 e^{\lambda(r)} P(r)}{c^4} + e^{\lambda(r)} - 1 \right) \;

Differential Equation of State for hydrostatic equilibrium:

\frac{dP(r)}{dr} = - \left( \frac{T_{00} g^{00} + T_{11} g^{11}}{2} \right) \frac{d\nu(r)}{dr} = - \left( \frac{\rho(r) c^2 + P(r)}{2} \right) \frac{d\nu(r)}{dr} \;

Where T_{00} and T_{11} are stress-energy tensor elements for hydrostatic density and hydrostatic pressure and g^{00} and g^{11} are inverse Einstein tensor metric elements, which are defined as:

g^{00} = \left( g_{00} \right)^{-1} = \left( e^{\nu(r)} \right)^{-1} = e^{-\nu(r)} \;
g^{11} = \left( g_{11} \right)^{-1} = \left( - e^{\lambda(r)} \right)^{-1} = - e^{-\lambda(r)} \;

This equation comes about by looking at the mixed tensor form of Einstein's Equations \frac{8 \pi G}{c^4} T^{\alpha}_{\beta} = G^{\alpha}_{\beta}.

Using the fact that the spatial components of the mixed tensor are equal, T^{r}_{r} = T^{\theta}_{\theta} = T^{\phi}_{\phi} = {\pm} P, the relations for T^{r}_{r} = T^{\theta}_{\theta} in terms of \nu',\lambda',\lambda, and r can be equated.

Rearranging terms and recognizing the expression for \frac{dP(r)}{dr} in terms of \nu',\lambda',\lambda,

and r yields the above expression.[2]

\frac{d\nu(r)}{dr} = - \frac{dP(r)}{dr} \left( \frac{2}{\rho(r) c^2 + P(r)} \right) = \frac{1}{r} \left( \frac{8 \pi G r^2 e^{\lambda(r)} P(r)}{c^4} + e^{\lambda(r)} - 1 \right) \;

Solve for P'(r):

Equation of State for hydrostatic equilibrium:

\frac{dP(r)}{dr} = -\frac{ \left( \rho(r) c^2 + P(r) \right) \left((e^{\lambda(r)} - 1)c^4 + 8 \pi G r^2 e^{\lambda(r)} P(r) \right)}{2 c^4 r} \;

Schwarzschild radius:

r_s = \frac{2 G M(r)}{c^2} \;

Metric identity:

e^{\lambda(r)} - 1 = \frac{r_s}{r - r_s} \;

Metric identity:

e^{\lambda(r)} = \left(1 - \frac{r_s}{r} \right)^{-1} \;

Integrating these metric identities via substitution results in the equation solution for the Equation of State for hydrostatic equilibrium.

Equation of State for hydrostatic equilibrium:

\frac{dP(r)}{dr} = - \frac{ \left( \rho(r) c^2 + P(r) \right) \left(c^4 r_s + 8 \pi G r^3 P(r) \right)}{2 c^4 r \left(r - r_s \right)} \;

Factoring out a 2 G c^2 from the numerator and multiplying c^2 through the denominator results in the Oppenheimer-Volkoff equation.

Oppenheimer-Volkoff equation:

\frac{dP(r)}{dr} = - \frac{G \left( \rho(r) c^2 + P(r) \right) \left(M(r) + 4 \pi r^3 \frac{P(r)}{c^2} \right)}{r \left(c^2 r - 2 G M(r) \right)} \;

Metric identity:

r (r - r_s) = r^2 \left( 1 - \frac{r_s}{r} \right) \;

Factoring out a c^2 from the numerator and integrating the metric identity via substitution results in the Tolman–Oppenheimer–Volkoff equation.

Tolman–Oppenheimer–Volkoff equation:

\frac{dP(r)}{dr} = - \frac{G}{r^2} \left( \rho(r) + \frac{P(r)}{c^2} \right) \left(M(r) + 4 \pi r^3 \frac{P(r)}{c^2} \right) \left( 1 - \frac{2 G M(r)}{c^2 r} \right)^{-1} \;

History

Richard C. Tolman analyzed spherically symmetric metrics in 1934 and 1939.[3][4] The form of the equation given here was derived by J. Robert Oppenheimer and George Volkoff in their 1939 paper, "On Massive Neutron Cores".[1] In this paper, the equation of state for a degenerate Fermi gas of neutrons was used to calculate an upper limit of ~0.7 solar masses for the gravitational mass of a neutron star. Since this equation of state is not realistic for a neutron star, this limiting mass is likewise incorrect. Modern estimates for this limit range from 1.5 to 3.0 solar masses.[5]

See also

References

  1. 1 2 3 4 5 J.R. Oppenheimer & G.M. Volkoff (1939). "On Massive Neutron Cores". Physical Review 55 (4): 374–381. Bibcode:1939PhRv...55..374O. doi:10.1103/PhysRev.55.374.
  2. Tolman, R. C., 1934, Relativity Thermodynamics and Cosmology(Oxford Press.) pp 243-244
  3. R.C. Tolman (1934). "Effect of Inhomogeneity on Cosmological Models". Proceedings of the National Academy of Sciences 20 (3): 169–176. Bibcode:1934PNAS...20..169T. doi:10.1073/pnas.20.3.169.
  4. R.C. Tolman (1939). "Static Solutions of Einstein's Field Equations for Spheres of Fluid". Physical Review 55 (4): 364–373. Bibcode:1939PhRv...55..364T. doi:10.1103/PhysRev.55.364.
  5. I. Bombaci (1996). "The Maximum Mass of a Neutron Star". Astronomy and Astrophysics 305: 871–877. Bibcode:1996A&A...305..871B.
This article is issued from Wikipedia - version of the Thursday, May 05, 2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.