Hamiltonian Form of Gravity around a Singularity
Abstract
We show that the Hamiltonian form of gravity in terms of connection and densitized triad variables admits a simpler structure close to a (spacelike) curvature singularity. To define this regime, we construct a limit to a vanishing (spatial) metric determinant. The cosmological and black hole spacetime solutions that characterize such an approach to a singularity are obtained. We also elucidate the connection of this generic limit to the ‘Carroll’ limit which could be interpreted as a special case.
I Introduction
A canonical quantization of full Hamiltonian gravity has turned out to be a difficult endeavour so far, given the complicated structure of the constraints. However, the answer to the question as to how Hamiltonian gravity could perceive the approach to a curvature singularity remains largely unknown. Here we address this issue, and demonstrate that such a regime could in fact be defined in terms of a simpler Hamiltonian structure.
The occurrance of a spacelike singularity in general relativity corresponds to the vanishing of the determinant () of the spatial metric. For a sufficiently well-behaved lapse, as assumed here, this limit is equivalent to the vanishing of the four-metric determinant (). Our goal here is to study the Hamiltonian structure of gravity theory based on the connection-triad formulation in this very limit, and to also analyze the resulting solutions.
First, we demonstrate how to implement the limit in the canonical theory of Hilbert-Palatini gravity, and find the set of truncated constraints. For convenience, we work directly with the time-gauge fixed theory, although the results are independent of any gauge-choice. Next, we apply this framework to specific examples of spacelike singularities. The limit is fairly general, in the sense that it could describe a local and anisotropic approach to singularity.
We also observe that under special assumptions, this limit admits a connection to the ‘Carroll’ (Levy Leblond-Sen Gupta) limit Lévy-Leblond (1965); Sen Gupta (1966); Teitelboim (1973); Henneaux (1979); Sengupta (2023), which may be defined for a speed of light that is too low () compared to another available velocity scale.
We conclude our analysis with a few relevant remarks in the final section.
II Canonical gravity in the limit
II.1 Constraints
To set the notation and variables, let us briefly recall the standard Hamiltonian representation of Hilbert-Palatini gravity. In terms of the tetrad and spin-connection fields , the Lagrangian density reads:
Here is the gravitational coupling and is the field strength.
The standard redefinition of the tetrad fields in terms of a scalar Lapse and the shift is given by Peldan (1994):
(1) |
The spatial metric is defined as: , and the tetrad determinant as: .
The resulting Hamiltonian density is given by a sum of constraints:
(2) |
where the constraints have the following expressions:
(3) |
In the above the momenta conjugate to the canonical coordinates are given by:
Next, we introduce a set of new variables as Barros e Sa (2001); Date et al. (2009); Kaul and Sengupta (2012):
(4) |
In these variables, the symplectric form reads:
Note that . In the above, the L.H.S. contains eighteen canonical pairs, whereas the R.H.S. has only twelve. Thus, the six redundant components () of correspond to vanishing momenta:
(5) |
A consistent time evolution of these leads to additional secondary constraints :
(6) |
Next, we fix the boost freedom () by choosing time-gauge . This in turn fixes the conjugate momenta as: . The constraints defined above then determine (and hence ) completely:
(7) |
The resulting constraints are displayed below:
(8) |
where in the last line are defined completely in terms of the densitized triad.
II.2 Limit to vanishing determinant
Let us now introduce the following scaling of the spatial triads:
(9) |
where the index in the scaling parameter (the round brackets indicate that these indices are not summed when repeated) runs over the three space coordinates. To emphasize, the scaling parameters along the spatial directions are allowed to be anisotropic and inhomogeneous in general. Using the requirement that the symplectic form (or, Poisson brackets) be preserved under the scaling (9), it acts on the phase space variables as:
(10) |
where . We define the limit to a vanishing triad determinant () as: for any . This implies that as the singularity is approached, the determinant vanishes faster than the rate of contraction or expansion of any of the three spatial directions.
The dependent variables and read:
where we have assumed the derivatives of the scaling parameters to satisfy the following properties:
(11) |
( above may or may not be equal).
Next, let us first rewrite the constraints in the rescaled variables and then take the limit. Note that the rotation constraints remain unaffected. Using (11), the diffeomorphism constraint (upto a rotation) is found to transform as:
Similarly, it is straightforward to check that the last two terms in the Hamiltonian constraint in eq.(II.1) are subleading, leading to a genuine truncation. Noting that the natural partial derivative on the rescaled variables is , as fixed by the transformation of the connection, the new set of constraints finally read:
(12) |
where the lapse and shift get rescaled as: . The Hamiltonian constraint is purely algebraic (upto an overall density factor) in the canonical variables. In other words, the non-polynomiality characteristic of the Hilbert-Palatini Hamiltonian constraint owing to the spatial scalar curvature term is absent here.
In this context, let us note that a Hamiltonian constraint of the same form as above may also be obtained in the first order form of ‘electric’ Carroll gravity associated with the limit Sengupta (2023). However, the latter limit corresponds to a constant scaling parameter acting on the variables. Hence, the limit is more general, and the limit might be interpreted as a special case with , as apparent from the scaling laws explicitly presented in ref.Sengupta (2023) 111We observe that a generalization of the ‘magnetic’ Carroll limit Bergshoeff et al. (2017); Henneaux and Salgado-Rebolledo (2021); Sengupta (2023) may be defined within the limiting formalism here, where the coordinates scale differently: . In this case, the Poisson brackets are not preserved. However, it is neither relevant nor our purpose here to discuss this limit any further..
III Examples: Approach to Spacelike singularity
III.1 Generalized Kasner Universe
As an example of an inhomogeneous and anisotropic universe, let us consider the following spacetime Kasner (1921); Belinsky et al. (1970); *bkl1:
Such a solution can exist in presence of matter in Einstein gravity. Assuming that the effect of matter could be neglected close to the spacelike curvature singularity at , we may use the Hamiltonian theory constructed in the previous section to analyze how the singularity is approached in this case.
The nontrivial components of the momenta are given by (suppressing the primes which are implicit in the variables and constraints):
The time evolution of the canonical pair with respect to the Hamiltonian constraint (II.2) leads to (in the gauge ):
(13) |
From the first equation above, the nontrivial components of the coordinates are obtained as:
The second evolution equation in (III.1) for the coordinates, along with the Hamiltonian constraint, imply:
(14) |
Assuming , these are solved as: . Next, the spatial diffeomorphism constraints lead to:
(15) |
These restrict the Kasner exponents as: , as the curvature singularity is approached. The rotation constraints are trivially satisfied.
To understand the behaviour of the derivatives ofthe scaling parameters, let us note that the approach to the singularity here could be parametrized as where ( being finite). Inserting this in the metric solution just obtained above, we find:
It is now straightforward to verify that the space derivatives satisfy all the conditions encoded in (11).
III.2 FLRW Cosmology with a scalar field
Here we consider the essential model underlying inflationary cosmology, namely, gravity coupled to a scalar field:
(16) |
The Hamiltonian theory in the time-gauge is now defined by the following set of constraints:
(17) |
While the canonical variables for gravity scale as in (10), the matter phase space pair do not transform. Rescaling the variables, taking the degenerate limit and then redefining the derivative and Lagrange multipliers exactly as earlier, we find once again that only the Hamitonian constraint gets modified (we suppress the primes as earlier):
(18) |
For an FLRW spacetime with a spatial curvature is given by:
the non-vanishing components of the canonical pairs are the following:
Using the above expressions, the Hamiltonian constraint and the time evolution of lead to the following equations of motion:
(19) |
For a comoving ideal fluid characterized by a density and pressure , this implies a stiff equation of state: (implying ). The spatial diffeomorphism constraints simplify as:
(20) |
Finally, the scalar equations of motion are given by:
(21) |
The solution to the above is: .
To emphasize, the limit rids the equations of motion of the spatial derivatives, the spatial curvature and the scalar potential. Hence, for any arbitrary scalar potential, it predicts the existence of a stiff phase right after the birth of the Universe (singularity).
III.3 Black hole
As our next example, we consider a spherically symmetric geometry where the radial coordinate is timelike:
(22) |
Proceeding as in the previous example, we obtain the only nontrivial components of the triad and connection as (in the gauge ):
The Hamiltonian constraint reads:
(23) |
whose solution is: , being an arbitrary function. All the remaining constraints are trivial, except the radial component of the diffomorphism:
(24) |
Finally, from the evolution equations for , we obtain:
(25) |
Comparing (24) and (25), we find the solution: , where . The solution for the metric (III.3) finally reads:
where the function has been absorbed (upto a const.) into a redefinition of the radial coordinate.
Note that the above is precisely the Schwarzschild metric in the limit , upon the identification , to the usual Schwarzschild coordinates and to the black hole mass. Thus, from the limiting theory, we find that the spacetime around the singularity could only be the interior Schwarzschild geometry when the radial coordinate is timelike.
IV Spherical symmetry
Next, for completeness, we explore the imports of the degenerate limit for a general geometry with spherical symmetry where the radial coordinate is spacelike:
The canonical variables defined by the Hamiltonian theory (II.2) are given by:
Here, the rotation and Hamiltonian constraints are trivial in this case. The only nontrivial constraint comes from the radial component of spatial diffeomorphism, implying:
(26) |
The lapse (or, ) remains arbitrary. Hence, the spherically symmetric solution in the limit is not necessarily static, contrary to Birkhoff’s theorem for vacuum Einstein gravity.
The result here could be translated to the electric Carroll limit as well, which represents a special case as emphasized earlier.
V Conclusions
Here, we construct a zero-determinant limit in canonical gravity theory and elucidate upon some of its attactive features. This limit, which encodes the approach to a curvature singularity, is shown to lead to a remarkably simpler Hamiltonian formulation of gravity. In particular, the non-polynomial terms arising from the spatial connection in the original constraints of Hilbert-Palatini gravity are absent in this limit.
From the Hamiltonian theory resulting from the limit, it follows that time derivatives dominate the spatial ones close to the singularity. This formulation may be contrasted with the BKL conjecture, which corresponds to a similar behaviour but a different limit in the Hamiltonian gravity in general Belinsky et al. (1970); *bkl1; Ashtekar et al. (2009).
In presence of scalar matter, we show that the limit predicts a stiff phase in the early universe. It is worth mentioning that such a phase had been first considered by Zeldovich Zel’dovich (1961) for dense nuclear matter, and later by Barrow Barrow (1978) in trying to explain the low anisotropy of the current universe in terms of an initial one with a low gravitational entropy. It is to be understood how such a phase could possibly affect any of the observable predictions of inflationary cosmology or its alternatives.
For the special case of constant scaling parameters, the limit here could be interpreted as equivalent to the ‘electric’ Carroll limit. Hence, all the results obtained based on the generic limit may be translated to the latter. There are two main imports of this connection. Firstly, we conclude that the ‘electric’ Carroll gravity is not really equivalent to the BKL limit, contrary to some earlier speculations in the literature. Further, the Birkhoff theorem as applicable to vacuum Einstein gravity for spherical symmetry is found not to get carried over to this limit when gravity dominates over matter.
Finally, let us emphasize that the limit, owing to the polynomial structure of the Hamiltonian constraint (upto a density factor), provides a rich prospect in the context of canonical quantum gravity. Any progress along these lines could be suggestive of hitherto unnoticed quantum features in the strong gravity regime. In this sense, the limit as constructed here should be useful even if a complete theoretical framework for gravity close to singularity is not available.
Acknowledgements.
This work is supported (in part) by the MATRICS project grant MTR/2021/000008, SERB, Govt. of India. Thanks are due to Ghanashyam Date, David Brizuela and Sayan Kar for their comments on this work, and to Alfredo Perez for a useful correspondence.References
- Lévy-Leblond (1965) J.-M. Lévy-Leblond, Une nouvelle limite non-relativiste du groupe de poincaré, Annales de l’I.H.P. Physique théorique 3, 1 (1965).
- Sen Gupta (1966) N. D. Sen Gupta, On an analogue of the Galilei group, Nuovo Cim. A 44, 512 (1966).
- Teitelboim (1973) C. Teitelboim, How commutators of constraints reflect the space-time structure, Annals Phys. 79, 542 (1973).
- Henneaux (1979) M. Henneaux, Geometry of Zero Signature Space-times, Bull. Soc. Math. Belg. 31, 47 (1979).
- Sengupta (2023) S. Sengupta, Hamiltonian form of Carroll gravity, Phys. Rev. D 107, 024010 (2023), arXiv:2208.02983 [gr-qc] .
- Peldan (1994) P. Peldan, Actions for gravity, with generalizations: A Review, Class. Quant. Grav. 11, 1087 (1994), arXiv:gr-qc/9305011 .
- Barros e Sa (2001) N. Barros e Sa, Hamiltonian analysis of general relativity with the Immirzi parameter, Int. J. Mod. Phys. D 10, 261 (2001), arXiv:gr-qc/0006013 .
- Date et al. (2009) G. Date, R. K. Kaul, and S. Sengupta, Topological Interpretation of Barbero-Immirzi Parameter, Phys. Rev. D 79, 044008 (2009), arXiv:0811.4496 [gr-qc] .
- Kaul and Sengupta (2012) R. K. Kaul and S. Sengupta, Topological parameters in gravity, Phys. Rev. D 85, 024026 (2012), arXiv:1106.3027 [gr-qc] .
- Bergshoeff et al. (2017) E. Bergshoeff, J. Gomis, B. Rollier, J. Rosseel, and T. ter Veldhuis, Carroll versus Galilei Gravity, JHEP 03, 165, arXiv:1701.06156 [hep-th] .
- Henneaux and Salgado-Rebolledo (2021) M. Henneaux and P. Salgado-Rebolledo, Carroll contractions of Lorentz-invariant theories, JHEP 11, 180, arXiv:2109.06708 [hep-th] .
- Kasner (1921) E. Kasner, Geometrical theorems on Einstein’s cosmological equations, Am. J. Math. 43, 217 (1921).
- Belinsky et al. (1970) V. A. Belinsky, I. M. Khalatnikov, and E. M. Lifshitz, Oscillatory approach to a singular point in the relativistic cosmology, Adv. Phys. 19, 525 (1970).
- Belinsky et al. (1982) V. a. Belinsky, I. m. Khalatnikov, and E. m. Lifshitz, A General Solution of the Einstein Equations with a Time Singularity, Adv. Phys. 31, 639 (1982).
- Ashtekar et al. (2009) A. Ashtekar, A. Henderson, and D. Sloan, Hamiltonian General Relativity and the Belinskii, Khalatnikov, Lifshitz Conjecture, Class. Quant. Grav. 26, 052001 (2009), arXiv:0811.4160 [gr-qc] .
- Zel’dovich (1961) Y. B. Zel’dovich, The equation of state at ultrahigh densities and its relativistic limitations, Zh. Eksp. Teor. Fiz. 41, 1609 (1961).
- Barrow (1978) J. D. Barrow, Quiescent cosmology, Nature 272, 211 (1978).