In general relativity, the Weyl metrics (named after the German-American mathematician Hermann Weyl)[1] are a class of static and axisymmetric solutions to Einstein's field equation. Three members in the renowned Kerr–Newman family solutions, namely the Schwarzschild, nonextremal Reissner–Nordström and extremal Reissner–Nordström metrics, can be identified as Weyl-type metrics.

Standard Weyl metrics edit

The Weyl class of solutions has the generic form[2][3]

 

(1)

where   and   are two metric potentials dependent on Weyl's canonical coordinates  . The coordinate system   serves best for symmetries of Weyl's spacetime (with two Killing vector fields being   and  ) and often acts like cylindrical coordinates,[2] but is incomplete when describing a black hole as   only cover the horizon and its exteriors.

Hence, to determine a static axisymmetric solution corresponding to a specific stress–energy tensor  , we just need to substitute the Weyl metric Eq(1) into Einstein's equation (with c=G=1):

 

(2)

and work out the two functions   and  .

Reduced field equations for electrovac Weyl solutions edit

One of the best investigated and most useful Weyl solutions is the electrovac case, where   comes from the existence of (Weyl-type) electromagnetic field (without matter and current flows). As we know, given the electromagnetic four-potential  , the anti-symmetric electromagnetic field   and the trace-free stress–energy tensor     will be respectively determined by

 

(3)
 

(4)

which respects the source-free covariant Maxwell equations:

 

(5.a)

Eq(5.a) can be simplified to:

 

(5.b)

in the calculations as  . Also, since   for electrovacuum, Eq(2) reduces to

 

(6)

Now, suppose the Weyl-type axisymmetric electrostatic potential is   (the component   is actually the electromagnetic scalar potential), and together with the Weyl metric Eq(1), Eqs(3)(4)(5)(6) imply that

 

(7.a)
 

(7.b)
 

(7.c)
 

(7.d)
 

(7.e)

where   yields Eq(7.a),   or   yields Eq(7.b),   or   yields Eq(7.c),   yields Eq(7.d), and Eq(5.b) yields Eq(7.e). Here   and   are respectively the Laplace and gradient operators. Moreover, if we suppose   in the sense of matter-geometry interplay and assume asymptotic flatness, we will find that Eqs(7.a-e) implies a characteristic relation that

 

(7.f)

Specifically in the simplest vacuum case with   and  , Eqs(7.a-7.e) reduce to[4]

 

(8.a)
 

(8.b)
 

(8.c)
 

(8.d)

We can firstly obtain   by solving Eq(8.b), and then integrate Eq(8.c) and Eq(8.d) for  . Practically, Eq(8.a) arising from   just works as a consistency relation or integrability condition.

Unlike the nonlinear Poisson's equation Eq(7.b), Eq(8.b) is the linear Laplace equation; that is to say, superposition of given vacuum solutions to Eq(8.b) is still a solution. This fact has a widely application, such as to analytically distort a Schwarzschild black hole.

We employed the axisymmetric Laplace and gradient operators to write Eqs(7.a-7.e) and Eqs(8.a-8.d) in a compact way, which is very useful in the derivation of the characteristic relation Eq(7.f). In the literature, Eqs(7.a-7.e) and Eqs(8.a-8.d) are often written in the following forms as well:


 

(A.1.a)
 

(A.1.b)
 

(A.1.c)
 

(A.1.d)
 

(A.1.e)

and

 

(A.2.a)
 

(A.2.b)
 

(A.2.c)
 

(A.2.d)

Considering the interplay between spacetime geometry and energy-matter distributions, it is natural to assume that in Eqs(7.a-7.e) the metric function   relates with the electrostatic scalar potential   via a function   (which means geometry depends on energy), and it follows that

 

(B.1)

Eq(B.1) immediately turns Eq(7.b) and Eq(7.e) respectively into

 

(B.2)
 

(B.3)

which give rise to

 

(B.4)

Now replace the variable   by  , and Eq(B.4) is simplified to

 

(B.5)

Direct quadrature of Eq(B.5) yields  , with   being integral constants. To resume asymptotic flatness at spatial infinity, we need   and  , so there should be  . Also, rewrite the constant   as   for mathematical convenience in subsequent calculations, and one finally obtains the characteristic relation implied by Eqs(7.a-7.e) that

 

(7.f)

This relation is important in linearize the Eqs(7.a-7.f) and superpose electrovac Weyl solutions.

Newtonian analogue of metric potential Ψ(ρ,z) edit

In Weyl's metric Eq(1),  ; thus in the approximation for weak field limit  , one has

 

(9)

and therefore

 

(10)

This is pretty analogous to the well-known approximate metric for static and weak gravitational fields generated by low-mass celestial bodies like the Sun and Earth,[5]

 

(11)

where   is the usual Newtonian potential satisfying Poisson's equation  , just like Eq(3.a) or Eq(4.a) for the Weyl metric potential  . The similarities between   and   inspire people to find out the Newtonian analogue of   when studying Weyl class of solutions; that is, to reproduce   nonrelativistically by certain type of Newtonian sources. The Newtonian analogue of   proves quite helpful in specifying particular Weyl-type solutions and extending existing Weyl-type solutions.[2]

Schwarzschild solution edit

The Weyl potentials generating Schwarzschild's metric as solutions to the vacuum equations Eq(8) are given by[2][3][4]

 

(12)

where

 

(13)

From the perspective of Newtonian analogue,   equals the gravitational potential produced by a rod of mass   and length   placed symmetrically on the  -axis; that is, by a line mass of uniform density   embedded the interval  . (Note: Based on this analogue, important extensions of the Schwarzschild metric have been developed, as discussed in ref.[2])

Given   and  , Weyl's metric Eq(1) becomes

 

(14)

and after substituting the following mutually consistent relations

 

(15)

one can obtain the common form of Schwarzschild metric in the usual   coordinates,

 

(16)

The metric Eq(14) cannot be directly transformed into Eq(16) by performing the standard cylindrical-spherical transformation  , because   is complete while   is incomplete. This is why we call   in Eq(1) as Weyl's canonical coordinates rather than cylindrical coordinates, although they have a lot in common; for example, the Laplacian   in Eq(7) is exactly the two-dimensional geometric Laplacian in cylindrical coordinates.

Nonextremal Reissner–Nordström solution edit

The Weyl potentials generating the nonextremal Reissner–Nordström solution ( ) as solutions to Eqs(7) are given by[2][3][4]

 

(17)

where

 

(18)

Thus, given   and  , Weyl's metric becomes

 

(19)

and employing the following transformations

 

(20)

one can obtain the common form of non-extremal Reissner–Nordström metric in the usual   coordinates,

 

(21)

Extremal Reissner–Nordström solution edit

The potentials generating the extremal Reissner–Nordström solution ( ) as solutions to Eqs(7) are given by[4] (Note: We treat the extremal solution separately because it is much more than the degenerate state of the nonextremal counterpart.)

 

(22)

Thus, the extremal Reissner–Nordström metric reads

 

(23)

and by substituting

 

(24)

we obtain the extremal Reissner–Nordström metric in the usual   coordinates,

 

(25)

Mathematically, the extremal Reissner–Nordström can be obtained by taking the limit   of the corresponding nonextremal equation, and in the meantime we need to use the L'Hospital rule sometimes.

Remarks: Weyl's metrics Eq(1) with the vanishing potential   (like the extremal Reissner–Nordström metric) constitute a special subclass which have only one metric potential   to be identified. Extending this subclass by canceling the restriction of axisymmetry, one obtains another useful class of solutions (still using Weyl's coordinates), namely the conformastatic metrics,[6][7]

 

(26)

where we use   in Eq(22) as the single metric function in place of   in Eq(1) to emphasize that they are different by axial symmetry ( -dependence).

Weyl vacuum solutions in spherical coordinates edit

Weyl's metric can also be expressed in spherical coordinates that

 

(27)

which equals Eq(1) via the coordinate transformation   (Note: As shown by Eqs(15)(21)(24), this transformation is not always applicable.) In the vacuum case, Eq(8.b) for   becomes

 

(28)

The asymptotically flat solutions to Eq(28) is[2]

 

(29)

where   represent Legendre polynomials, and   are multipole coefficients. The other metric potential  is given by[2]

 

(30)

See also edit

References edit

  1. ^ Weyl, H., "Zur Gravitationstheorie," Ann. der Physik 54 (1917), 117–145.
  2. ^ a b c d e f g h Jeremy Bransom Griffiths, Jiri Podolsky. Exact Space-Times in Einstein's General Relativity. Cambridge: Cambridge University Press, 2009. Chapter 10.
  3. ^ a b c Hans Stephani, Dietrich Kramer, Malcolm MacCallum, Cornelius Hoenselaers, Eduard Herlt. Exact Solutions of Einstein's Field Equations. Cambridge: Cambridge University Press, 2003. Chapter 20.
  4. ^ a b c d R Gautreau, R B Hoffman, A Armenti. Static multiparticle systems in general relativity. IL NUOVO CIMENTO B, 1972, 7(1): 71-98.
  5. ^ James B Hartle. Gravity: An Introduction To Einstein's General Relativity. San Francisco: Addison Wesley, 2003. Eq(6.20) transformed into Lorentzian cylindrical coordinates
  6. ^ Guillermo A Gonzalez, Antonio C Gutierrez-Pineres, Paolo A Ospina. Finite axisymmetric charged dust disks in conformastatic spacetimes. Physical Review D, 2008, 78(6): 064058. arXiv:0806.4285v1
  7. ^ Antonio C Gutierrez-Pineres, Guillermo A Gonzalez, Hernando Quevedo. Conformastatic disk-haloes in Einstein-Maxwell gravity. Physical Review D, 2013, 87(4): 044010. [1]