Tetradic Palatini action – Wikipedia

The Einstein–Hilbert action for general relativity was first formulated purely in terms of the space-time metric. To take the metric and affine connection as independent variables in the action principle was first considered by Palatini.[1] It is called a first order formulation as the variables to vary over involve only up to first derivatives in the action and so doesn’t overcomplicate the Euler–Lagrange equations with higher derivative terms. The tetradic Palatini action is another first-order formulation of the Einstein–Hilbert action in terms of a different pair of independent variables, known as frame fields and the spin connection. The use of frame fields and spin connections are essential in the formulation of a generally covariant fermionic action (see the article spin connection for more discussion of this) which couples fermions to gravity when added to the tetradic Palatini action.

Not only is this needed to couple fermions to gravity and makes the tetradic action somehow more fundamental to the metric version, the Palatini action is also a stepping stone to more interesting actions like the self-dual Palatini action which can be seen as the Lagrangian basis for Ashtekar’s formulation of canonical gravity (see Ashtekar’s variables) or the Holst action which is the basis of the real variables version of Ashtekar’s theory. Another important action is the Plebanski action (see the entry on the Barrett–Crane model), and proving that it gives general relativity under certain conditions involves showing it reduces to the Palatini action under these conditions.

Here we present definitions and calculate Einstein’s equations from the Palatini action in detail. These calculations can be easily modified for the self-dual Palatini action and the Holst action.

Some definitions[edit]

We first need to introduce the notion of tetrads. A tetrad is an orthonormal vector basis in terms of which the space-time metric looks locally flat,

where

ηIJ=diag(1,1,1,1){displaystyle eta _{IJ}={text{diag}}(-1,1,1,1)}

is the Minkowski metric. The tetrads encode the information about the space-time metric and will be taken as one of the independent variables in the action principle.

Now if one is going to operate on objects that have internal indices one needs to introduce an appropriate derivative (covariant derivative). We introduce an arbitrary covariant derivative via

Where

ωαIJ{displaystyle {omega _{alpha I}}^{J}}

is a spin (Lorentz) connection one-form (the derivative annihilates the Minkowski metric

ηIJ{displaystyle eta _{IJ}}

). We define a curvature via

We obtain

We introduce the covariant derivative which annihilates the tetrad,

The connection is completely determined by the tetrad. The action of this on the generalized tensor

VβI{displaystyle V_{beta }^{I}}

is given by

We define a curvature

RαβIJ{displaystyle {R_{alpha beta }}^{IJ}}

by

This is easily related to the usual curvature defined by

via substituting

Vγ=VIeγI{displaystyle V_{gamma }=V_{I}e_{gamma }^{I}}

into this expression (see below for details). One obtains,

for the Riemann tensor, Ricci tensor and Ricci scalar respectively.

The tetradic Palatini action[edit]

The Ricci scalar of this curvature can be expressed as

eIαeJβΩαβIJ.{displaystyle e_{I}^{alpha }e_{J}^{beta }{Omega _{alpha beta }}^{IJ}.}

The action can be written

where

e=g{displaystyle e={sqrt {-g}}}

but now

g{displaystyle g}

is a function of the frame field.

We will derive the Einstein equations by varying this action with respect to the tetrad and spin connection as independent quantities.

As a shortcut to performing the calculation we introduce a connection compatible with the tetrad,

αeβI=0.{displaystyle nabla _{alpha }e_{beta }^{I}=0.}

[2] The connection associated with this covariant derivative is completely determined by the tetrad. The difference between the two connections we have introduced is a field

CαIJ{displaystyle {C_{alpha I}}^{J}}

defined by

We can compute the difference between the curvatures of these two covariant derivatives (see below for details),

The reason for this intermediate calculation is that it is easier to compute the variation by reexpressing the action in terms of

{displaystyle nabla }

and

CαIJ{displaystyle {C_{alpha }}^{IJ}}

and noting that the variation with respect to

ωαIJ{displaystyle {omega _{alpha }}^{IJ}}

is the same as the variation with respect to

CαIJ{displaystyle {C_{alpha }}^{IJ}}

(when keeping the tetrad fixed). The action becomes

We first vary with respect to

CαIJ{displaystyle {C_{alpha }}^{IJ}}

. The first term does not depend on

CαIJ{displaystyle {C_{alpha }}^{IJ}}

so it does not contribute. The second term is a total derivative. The last term yields

We show below that this implies that

CαIJ=0{displaystyle {C_{alpha }}^{IJ}=0}

as the prefactor

eM[aeNb]δ[IMδJ]K{displaystyle e_{M}^{[a}e_{N}^{b]}delta _{[I}^{M}delta _{J]}^{K}}

is non-degenerate. This tells us that

{displaystyle nabla }

coincides with

D{displaystyle D}

when acting on objects with only internal indices. Thus the connection

D{displaystyle D}

is completely determined by the tetrad and

Ω{displaystyle Omega }

coincides with

R{displaystyle R}

. To compute the variation with respect to the tetrad we need the variation of

e=deteαI{displaystyle e=det e_{alpha }^{I}}

. From the standard formula

we have

δe=eeIαδeαI{displaystyle delta e=ee_{I}^{alpha }delta e_{alpha }^{I}}

. Or upon using

δ(eαIeIα)=0{displaystyle delta left(e_{alpha }^{I}e_{I}^{alpha }right)=0}

, this becomes

δe=eeαIδeIα{displaystyle delta e=-ee_{alpha }^{I}delta e_{I}^{alpha }}

. We compute the second equation by varying with respect to the tetrad,

One gets, after substituting

ΩαβIJ{displaystyle {Omega _{alpha beta }}^{IJ}}

for

RαβIJ{displaystyle {R_{alpha beta }}^{IJ}}

as given by the previous equation of motion,

which, after multiplication by

eIβ{displaystyle e_{Ibeta }}

just tells us that the Einstein tensor

Rαβ12Rgαβ{displaystyle R_{alpha beta }-{tfrac {1}{2}}Rg_{alpha beta }}

of the metric defined by the tetrads vanishes. We have therefore proved that the Palatini variation of the action in tetradic form yields the usual Einstein equations.

Generalizations of the Palatini action[edit]

We change the action by adding a term

This modifies the Palatini action to

where

This action given above is the Holst action, introduced by Holst[3] and

γ{displaystyle gamma }

is the Barbero-Immirzi parameter whose role was recognized by Barbero[4] and Immirizi.[5] The self dual formulation corresponds to the choice

γ=i{displaystyle gamma =-i}

.

It is easy to show these actions give the same equations. However, the case corresponding to

γ=±i{displaystyle gamma =pm i}

must be done separately (see article self-dual Palatini action). Assume

γ±i{displaystyle gamma not =pm i}

, then

PIJMN{displaystyle {P^{IJ}}_{MN}}

has an inverse given by

(note this diverges for

γ=±i{displaystyle gamma =pm i}

). As this inverse exists the generalization of the prefactor

eM[aeNb]δ[IMδJ]K{displaystyle e_{M}^{[a}e_{N}^{b]}delta _{[I}^{M}delta _{J]}^{K}}

will also be non-degenerate and as such equivalent conditions are obtained from variation with respect to the connection. We again obtain

CαIJ=0{displaystyle {C_{alpha }}^{IJ}=0}

. While variation with respect to the tetrad yields Einstein’s equation plus an additional term. However, this extra term vanishes by symmetries of the Riemann tensor.

Details of calculation[edit]

Relating usual curvature to the mixed index curvature[edit]

The usual Riemann curvature tensor

Rαβγδ{displaystyle {R_{alpha beta gamma }}^{delta }}

is defined by

To find the relation to the mixed index curvature tensor let us substitute

Vγ=eγIVI{displaystyle V_{gamma }=e_{gamma }^{I}V_{I}}

where we have used

αeβI=0{displaystyle nabla _{alpha }e_{beta }^{I}=0}

. Since this is true for all

Vδ{displaystyle V_{delta }}

we obtain

Using this expression we find

Contracting over

α{displaystyle alpha }

and

β{displaystyle beta }

allows us write the Ricci scalar

Difference between curvatures[edit]

The derivative defined by

DαVI{displaystyle D_{alpha }V_{I}}

only knows how to act on internal indices. However, we find it convenient to consider a torsion-free extension to spacetime indices. All calculations will be independent of this choice of extension. Applying

Da{displaystyle {mathcal {D}}_{a}}

twice on

VI{displaystyle V_{I}}

,

where

Γ¯αβγ{displaystyle {overline {Gamma }}_{alpha beta }^{gamma }}

is unimportant, we need only note that it is symmetric in

α{displaystyle alpha }

and

β{displaystyle beta }

as it is torsion-free. Then

Hence:

Varying the action with respect to the field

We would expect

a{displaystyle nabla _{a}}

to also annihilate the Minkowski metric

ηIJ=eβIeJβ{displaystyle eta _{IJ}=e_{beta I}e_{J}^{beta }}

. If we also assume that the covariant derivative

Dα{displaystyle {mathcal {D}}_{alpha }}

annihilates the Minkowski metric (then said to be torsion-free) we have,

Implying

From the last term of the action we have from varying with respect to

CαIJ,{displaystyle {C_{alpha I}}^{J},}

δSEH=δd4xeeMγeNβC[γMKCβ]KN=δd4xeeM[γeNβ]CγMKCβKN=δd4xeeM[γeNβ]CγMKCβKN=d4xeeM[γeNβ](δγαδMIδJKCβKN+CγMKδβαδKIδJN)δCαIJ=d4xe(eI[αeNβ]CβJN+eM[βeJα]CβMI)δCαIJ{displaystyle {begin{aligned}delta S_{EH}&=delta int d^{4}x;e;e_{M}^{gamma }e_{N}^{beta }{C_{[gamma }}^{MK}{C_{beta ]K}}^{N}\&=delta int d^{4}x;e;e_{M}^{[gamma }e_{N}^{beta ]}{C_{gamma }}^{MK}{C_{beta K}}^{N}\&=delta int d^{4}x;e;e^{M[gamma }e_{N}^{beta ]}{C_{gamma M}}^{K}{C_{beta K}}^{N}\&=int d^{4}x;ee^{M[gamma }e_{N}^{beta ]}left(delta _{gamma }^{alpha }delta _{M}^{I}delta _{J}^{K}{C_{beta K}}^{N}+{C_{gamma M}}^{K}delta _{beta }^{alpha }delta _{K}^{I}delta _{J}^{N}right)delta {C_{alpha I}}^{J}\&=int d^{4}x;eleft(e^{I[alpha }e_{N}^{beta ]}{C_{beta J}}^{N}+e^{M[beta }e_{J}^{alpha ]}{C_{beta M}}^{I}right)delta {C_{alpha I}}^{J}end{aligned}}}

or

or

where we have used

CβKI=CβIK{displaystyle C_{beta KI}=-C_{beta IK}}

. This can be written more compactly as

Vanishing of

We will show following the reference “Geometrodynamics vs. Connection Dynamics”[6] that

implies

CαIJ=0.{displaystyle {C_{alpha I}}^{J}=0.}

First we define the spacetime tensor field by

Then the condition

CαIJ=Cα[IJ]{displaystyle C_{alpha IJ}=C_{alpha [IJ]}}

is equivalent to

Sαβγ=Sα[βγ]{displaystyle S_{alpha beta gamma }=S_{alpha [beta gamma ]}}

. Contracting Eq. 1 with

eαIeγJ{displaystyle e_{alpha }^{I}e_{gamma }^{J}}

one calculates that

As

Sαβγ=CαIJeβIeJγ,{displaystyle {S_{alpha beta }}^{gamma }={C_{alpha I}}^{J}e_{beta }^{I}e_{J}^{gamma },}

we have

Sβγβ=0.{displaystyle {S_{beta gamma }}^{beta }=0.}

We write it as

and as

eαI{displaystyle e_{alpha }^{I}}

are invertible this implies

Thus the terms

CβIKeKβeJα,{displaystyle {C_{beta I}}^{K}e_{K}^{beta }e_{J}^{alpha },}

and

CβJKeIαeKβ{displaystyle {C_{beta J}}^{K}e_{I}^{alpha }e_{K}^{beta }}

of Eq. 1 both vanish and Eq. 1 reduces to

If we now contract this with

eγIeδJ{displaystyle e_{gamma }^{I}e_{delta }^{J}}

, we get

or

Since we have

Sαβγ=Sα[βγ]{displaystyle S_{alpha beta gamma }=S_{alpha [beta gamma ]}}

and

Sαβγ=S(αβ)γ{displaystyle S_{alpha beta gamma }=S_{(alpha beta )gamma }}

, we can successively interchange the first two and then last two indices with appropriate sign change each time to obtain,

Implying

or

and since the

eαI{displaystyle e_{alpha }^{I}}

are invertible, we get

CαIJ=0{displaystyle C_{alpha IJ}=0}

. This is the desired result.

See also[edit]

References[edit]

  1. ^ A. Palatini (1919) Deduzione invariantiva delle equazioni gravitazionali dal principio di Hamilton, Rend. Circ. Mat. Palermo 43, 203-212 [English translation by R.Hojman and C. Mukku in P.G. Bergmann and V. De Sabbata (eds.) Cosmology and Gravitation, Plenum Press, New York (1980)]
  2. ^ A. Ashtekar “Lectures on non-perturbative canonical gravity” (with invited contributions), Bibliopolis, Naples 19988.
  3. ^ Holst, Sören (1996-05-15). “Barbero’s Hamiltonian derived from a generalized Hilbert-Palatini action”. Physical Review D. 53 (10): 5966–5969. arXiv:gr-qc/9511026. Bibcode:1996PhRvD..53.5966H. doi:10.1103/physrevd.53.5966. ISSN 0556-2821. PMID 10019884. S2CID 15959938.
  4. ^ Barbero G., J. Fernando (1995-05-15). “Real Ashtekar variables for Lorentzian signature space-times”. Physical Review D. 51 (10): 5507–5510. arXiv:gr-qc/9410014. Bibcode:1995PhRvD..51.5507B. doi:10.1103/physrevd.51.5507. ISSN 0556-2821. PMID 10018309. S2CID 16314220.
  5. ^ Immirzi, Giorgio (1997-10-01). “Real and complex connections for canonical gravity”. Classical and Quantum Gravity. IOP Publishing. 14 (10): L177–L181. arXiv:gr-qc/9612030. doi:10.1088/0264-9381/14/10/002. ISSN 0264-9381. S2CID 5795181.
  6. ^ Romano, Joseph D. (1993). “Geometrodynamics vs. connection dynamics”. General Relativity and Gravitation. 25 (8): 759–854. arXiv:gr-qc/9303032. Bibcode:1993GReGr..25..759R. doi:10.1007/bf00758384. ISSN 0001-7701. S2CID 119359223.