Next Article in Journal
Combining Electromyography and Tactile Myography to Improve Hand and Wrist Activity Detection in Prostheses
Previous Article in Journal
Estimating Potential Methane Emission from Municipal Solid Waste and a Site Suitability Analysis of Existing Landfills in Delhi, India
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Non-Euclidean Hydrodynamic Klein–Gordon Equation with Perturbative Self-Interacting Field

1
National Council of Research of Italy, Area of Pisa, Moruzzi 1, 56124 Pisa, Italy
2
Interdepartmental Center “E.Piaggio”, University of Pisa, 56124 Pisa, Italy
Technologies 2017, 5(4), 63; https://doi.org/10.3390/technologies5040063
Submission received: 31 July 2017 / Revised: 22 August 2017 / Accepted: 26 September 2017 / Published: 4 October 2017
(This article belongs to the Section Quantum Technologies)

Abstract

:
In this paper the quantum hydrodynamic approach for the Klein–Gordon equation (KGE) owning a perturbative self-interaction term is developed. The generalized model to non-Euclidean space–time allows for the determination of the quantum energy impulse tensor density of mesons, for the gravitational equation of quantum mechanical systems.

1. Introduction

A longstanding objective of modern physics is to describe the quantum mechanics and the quantum fields in the framework of general relativity. On an ordinary scale, the two theories are decoupled and even if theoretical models are available [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16], it not easy to find experimental confirmations.
A first physical evidence of quantum gravity (QG) comes by way of the corrections to the Newton law, derived from the Galilean limit of the theory [17,18,19,20,21,22,23,24,25,26,27,28,29]. The Hawking radiation of a black hole can also be justified in the frame of QG [30,31,32,33,34,35,36,37,38].
Moreover, given that the uncertainty principle forbids a black hole (BH) collapse [39,40,41,42,43,44] and that the quantization of a black hole requires the existence of a fundamental state with a minimum mass, the BH cannot have a mass smaller than that of Planck and comparable with those of elementary particles [45].
Another phenomenon connected to QG is the cosmological constant that it is needed to model the motion of galaxies on the universal scale [46,47,48,49,50,51,52,53,54].
If we try to define the QG equation from the minimum action principle (MAP), the quantizing action is not self-contained under renormalization, since loop diagrams generate terms not present in the initial expression [55,56,57,58,59].
In the present work the author makes use of the hydrodynamic quantum description (HQD) [60,61,62,63,64,65,66,67,68] that has the advantage of being embedded in a classical-like formalism, where the quantization is implicitly contained into the equation of evolution [68], for deriving a gravity equation for quantum mechanical systems.
The quantum hydrodynamic approach, firstly developed by Madelung [62], describes the evolution of the complex wavefunction ψ = | ψ | exp i S as a function of the two real variables, | ψ | and S [60,62,63,69,70,71,72,73]. The model gives rise to classical-like analogy describing the motion of particles’ density n = | ψ | 2 owning the 4-impulse p μ = μ S .
As shown by Weiner et al. [71], the outputs of the quantum hydrodynamic model agree with the outputs of the Schrödinger problem, but not only for the semi-classical limit or for a single particle [62,63,69]. More recently, Koide and Kodama [72], showed that it agrees with the outputs of the stochastic variational method.
Recently, the author of this article has shown that the hydrodynamic approach can be derived from the properties of vacuum on a small scale [73].
Moreover, as shown by Bohm and Hiley [74,75] the hydrodynamic approach can be generalized for the description of the quantum fields.
The present work develops the quantum hydrodynamic form of the Klein–Gordon equation (KGE) containing an additional perturbative self-interaction term.
The interest of such a description lies in the fact that this kind of KGE can describe the states of bosons, such as mesons. The goal of this paper is to obtain the energy–impulse tensor of such a particle, for the coupling with the Einstein gravitational equation (GE) for quantum mechanical systems [68].
As shown in study [68], the quantum mechanical QG defines the gravitational coupling of classical fields and the commutation rules in curved space to be used for their quantization.
The paper is organized as follows: in Section 2, the hydrodynamic KGE with a perturbative self-interaction term is derived for scalar particle and for a charged boson. In Section 3, the formulae are generalized to a non-Euclidean space–time.

2. The Hydrodynamic KGE with Perturbative Self-Interaction

In this section, the Euclidean hydrodynamic representation of the KGE is derived for a scalar uncharged boson with a perturbative self-interaction term that reads
μ μ ψ = m 2 c 2 2 ψ λ | ψ | 3
Following the procedure given in studies [45,68] the hydrodynamic equations of motion are given by the Hamilton–Jacobi type equation
g μ ν μ S   ν S 2 ( μ μ | ψ | | ψ | + λ | ψ | 2 ) m 2 c 2 = 0
coupled with the current equation [67,69]
μ ( | ψ | 2 μ S ) = 0
where
S = 2 i ln [ ψ ψ * ]
and where
| ψ | 2 m μ S = J μ = i 2 m ( ψ * μ ψ ψ μ ψ * )
where the signature (1, −1, −1, −1) has been used.
Moreover, being the 4-impulse of the hydrodynamic analogy
p μ = μ S
it follows that
J μ = ( c ρ , J i ) = | ψ | 2 p μ m
where
ρ = | ψ | 2 m c 2 S t
Moreover, by using (6), Equation (2) can be rewritten as
μ S μ S = p μ p μ = ( E 2 c 2 p 2 ) = m 2 c 2 ( 1 V q u m c 2 )
where
V q u = 2 m ( μ μ | ψ | | ψ | + λ | ψ | 2 ) ,
and where p 2 = p i p i is the modulus of the spatial 3-impulse.
As shown in studies [67,68], given the hydrodynamic Lagrangean function
L = d S d t = S t + S q i q ˙ i = p μ q ˙ μ = 1 2 n a n | ψ n | exp [ i S n ] ( i q ˙ μ μ ln | ψ n | + L n ) n a n | ψ n | exp [ i S n ] 1 2 n a * n | ψ n | exp [ i S n ] ( i q ˙ μ μ ln | ψ n | L n ) n a * n | ψ n | exp [ i S n ]
Equation (2) can be expressed by the following system of Lagrangean equations of motion
p μ = L q ˙ μ ,
p ˙ μ = L q μ .
that for the eigenstates read
p n μ = L n q ˙ μ ,
p ˙ n μ = L n q μ
where
L n = ( ± ) m c 2 γ 1 V q u ( n ) m c 2 = ( ± ) m c 2 γ 1 + 2 m 2 c 2 ( μ μ | ψ n | | ψ n | + λ | ψ n | 2 )
where c is the speed of light and γ = 1 1 q ˙ 2 c 2 . Generally speaking, for eigenstates, for which it holds E = E n = c o n s t , it follows that
( E n 2 c 2 p n 2 ) = m 2 c 2 ( 1 V q u ( n ) m c 2 ) = m 2 γ 2   c 2 ( 1 V q u ( n ) m c 2 ) m 2 γ 2   q ˙ 2 ( 1 V q u ( n ) m c 2 )
from where it follows that
E n = ± m γ c 2 1 V q u ( n ) m c 2 = ± m γ c 2 1 + 2 m 2 c 2 ( μ μ | ψ n | | ψ n | + λ | ψ n | 2 )
(where the minus sign stands for antiparticles) and, by using (17), that
p n μ = ± m γ   q ˙ μ 1 V q u ( n ) m c 2 = E n c 2 q ˙ μ .
Following the hydrodynamic protocol [68], the eigenstates are represented by the stationary solutions of the hydrodynamic equations of motion obtained by deriving p μ ( q ˙ , q ) from (14) and then inserting it into (15) that leads to
d p n μ d s = γ c L n q μ = ± d d s ( m c u μ ( 1 V q u ) ( n ) m c 2 ) ) = ± m c q μ 1 V q u ( n ) m c 2
where
u μ = γ c q ˙ μ ,
and to
± m c 1 V q u ( n ) m c 2 d u μ d s = ( ± ) m c u μ d d s ( 1 V q u ( n ) m c 2 ) ± m c q μ ( 1 V q u ( n ) m c 2 ) = γ c T n μ ν q ν
where for eigenstates, the quantum energy-impulse tensor (QEIT) T n μ ν reads [45,68]
T n μ ν = ( q ˙ μ L n q ˙ ν L n δ μ ν ) = ± m c 2 γ 1 V q u ( n ) m c 2 ( u μ u ν δ μ ν ) .
leading to the quantum energy impulse tensor density (QEITD) [45,68]
T n μ ν = q ˙ μ L n n q ˙ ν L n n δ μ ν = | ψ n | 2 ( q ˙ μ L n q ˙ ν L n δ μ ν ) = | ψ n | 2 T n μ ν
where L = | ψ | 2 L is the (hydrodynamic) Lagrangian density and L is the hydrodynamic Lagrangian function.
The quantization condition is brought inside the above quantum hydrodynamic relations by the quantum potential V q u ( n ) contained in the term 1 V q u ( n ) m c 2 . The analog quantities of the classical limit are recovered for 0 , and the energy–impulse tensor (23) in Equation (22) leads to the classical equation of motion.
Moreover, by using the identity
S n = 2 i ln [ ψ n ψ n * ] ,
the QEITD (24) can be written as a function of the wave function as following
T n μ ν = ± | ψ | 2 c 2 ( S n t ) 1 ( p μ p ν p α p α δ μ ν ) = ± m | ψ n | 2 c 2 ( 2 i m 2 c 2 ln [ ψ n ψ * n ] t ) 1 ( ( 2 m c ) 2 ln [ ψ n ψ * n ] q μ ln [ ψ n ψ * n ] q ν + ( 1 V q u ( n ) m c 2 ) δ μ ν )

Charged Boson

In the case of a charged boson, Equations (1)–(3) read, respectively,
D μ D μ ψ = ψ ( m 2 c 2 2 + λ | ψ | 2 )
where D μ = μ e i A μ ,
( μ S + e A μ ) ( μ S + e A μ ) = m 2 c 2 + 2 ( μ μ | ψ | | ψ | + λ | ψ | 2 )
J μ q μ = 0
where the 4-current J μ reads
J μ = ( c ρ , J i ) = 2 i m ( ψ * ( μ e i A μ ) ψ ψ ( μ + e i A μ ) ψ * ) = | ψ | 2 m [ p μ e A μ ] = | ψ | 2 m π μ
and
μ S = p μ = π μ + e A μ = ( E c , p i )
(where π μ is the mechanical momentum) [67,68] and where
ρ = | ψ | 2 m c 2 [ S t + e ϕ ] .
Moreover, analogously to (9) and (17)–(19), from (28) it follows that
π n μ = 1 c 2 [ S n t e ϕ ] q ˙ μ = E n e ϕ c 2 q ˙ μ = p n μ e A μ
that leads to
E n e ϕ = ± m γ c 2 1 V q u ( n ) m c 2 = ± m γ c 2 1 + 2 m 2 c 2 ( μ μ | ψ n | | ψ n | + λ | ψ n | 2 ) ,
to
p n μ e A μ = ± m γ   q ˙ μ 1 V q u ( n ) m c 2 = E n c 2 q ˙ μ
and to
L n = d S n d t = S n t + S n q i q ˙ i = p n μ q ˙ μ = ( 1 c 2 S n t + e c 2 ϕ ) 1 p n μ ( p n μ e A μ ) = ( ± ) ( m c 2 γ 1 + 2 m 2 c 2 ( μ μ | ψ n | | ψ n | + λ | ψ n | 2 ) e A μ q ˙ μ )
that by using (25), as a function of ψ and A μ , reads
L n = i 2 c 2 ( ln [ ψ n ψ n * ] t + 2 i e ϕ ) 1 ln [ ψ n ψ n * ] q μ ( ln [ ψ n ψ n * ] q μ 2 i e A μ ) .
For = 0 the Lagrangean (37) acquires the known classical expression L = ( m c 2 γ + e A μ q ˙ μ ) . Moreover, with the help of (25), (30), (33)–(35) it follows that
T n μ ν   = ( ± ) m | ψ n | 2 c 2 γ 1 V q u ( n ) m c 2 ( ( u μ + ( 1 V q u ( n ) m c 2 ) 1 e m c A μ ) u ν ( 1 + ( 1 V q u ( n ) m c 2 ) 1 e m c A α u α ) δ μ ν )
that by using (25), (30) and (35) we can be expressed as a function of the wave function as
T μ ν = | ψ n | 2 c 2 2 i ( t ln [ ψ n ψ n * ] 2 i e ϕ ) 1 ( ( ln [ ψ n ψ n * ] q μ 2 i e A μ ) ln [ ψ n ψ n * ] q ν ( ln [ ψ n ψ n * ] q α 2 i e A α ) ln [ ψ n ψ n * ] q α δ μ ν )
The above equations are coupled to the Maxwell equation:
F μ ν ; ν = 4 π   J μ
where [75]
F μ ν = ( A ν , μ A μ , ν ) = ( μ A ν ν A μ ) ,
and where
A μ = ( ϕ c ,   A i )
is the potential 4-vector.

3. Non-Euclidean Space–Time

The quartic self-interaction is introduced in the KGE in order to describe the states of charged ( ± 1 ) bosons (e.g., mesons) [76]. The importance of having the hydrodynamic description of bosons [68] lies in the fact that it becomes possible to derive its quantum energy–impulse tensor that can couple them to the Einstein gravitational equation for quantum mechanical systems [68].
By using the General Physics Covariance postulate [3,68], it is possible to derive the non-Euclidean expression of the hydrodynamic model of the KGE:
( μ ψ ) , μ = 1 g μ ( g g μ ν ν ψ ) = ψ ( m 2 c 2 2 + λ | ψ | 2 )
Equations (2) and (3) in a non-Euclidean space–time read, respectively,
g μ ν μ S   ν S + m V q u m 2 c 2 = 0 1 g μ g ( g μ ν | ψ | 2 ν S ) = 0
where
V q u = 2 m ( 1 | ψ | g μ g ( g μ ν ν | ψ | ) + λ | ψ | 2 ) .
Moreover, by using the Lagrangean function
L n = d S n d t = g μ ν   p n ν q ˙ μ ,
the covariant form of the motion Equations (14) and (15) reads
p n μ = L n q ˙ μ
D t p n μ = L n q μ ,
where
D t p μ = p ˙ μ Γ μ λ ν p ν q ˙ λ
is the total covariant derivative respect the time and where Γ μ λ ν are the Christoffel symbols.
Equations (47) and (48) leads to the motion equation
D t ( L n q ˙ μ ) = L n q μ
where L n reads
L n = ( ± ) m c 2 γ 1 V q u ( n ) m c 2 = ( ± ) m c 2 g μ ν q ˙ ν q ˙ μ c 2 1 V q u ( n ) m c 2 .
From (50) it follows that the motion equation, describing the evolution of the particle density n = | ψ n | 2 moving with the hydrodynamic impulse (6), reads
d u μ d t 1 2 c γ g λ κ q μ u λ u κ = c γ ln 1 V q u ( n ) m c 2 q μ + c γ ln γ q μ u μ d d t ln 1 V q u m c 2
where the stationary condition d u μ d t = 0 , that determines the balance between the “force” of gravity and that one of the quantum potential, defines the stationary equation for the eigenstates
1 2 c γ μ g λ κ u λ u κ + c γ μ ln γ = c γ μ ln 1 V q u ( n ) m c 2 + u μ d d t ln 1 V q u m c 2 .
where 1 g = | g ν μ | = J a c 2 , where J a c is Jacobean of the transformation of the Galilean co-ordinates to non-Euclidean ones and where g ν μ is the metric tensor defined by the gravitational equation for quantum mechanical systems [68,75]
R μ ν 1 2 g μ ν R α α Λ g μ ν = 8 π G c 4 T μ ν  
where the quantum energy impulse tensor density reads
T ( n ) μ ν = ± | ψ n | 2 m c 2 g μ ν q ˙ ν q ˙ μ c 2 1 V q u ( n ) m c 2 ( u μ u ν g α β u α u β g μ ν ) .
And where the cosmological energy-impulse density Λ [68], for eigenstates, reads
Λ = 8 π G c 4 | ψ n | 2 L 0 = 8 π G c 4 m | ψ n | 2 c 2 γ
where for scalar uncharged particles leads to
L 0 = lim 0 L = ( ± ) m c 2 γ .
The quantum energy impulse tensor density for the particle density distribution | ψ n | 2 owing with the hydrodynamic impulse (6), as a function of the wave function reads
T μ ν = T μ α g α ν = | ψ | 2 m 2 c 4 ( S t ) 1 ( p μ p ν m 2 c 2 ( 1 V q u m c 2 ) g μ ν ) = m | ψ | 2 c 2 ( 2 i m 2 c 2 ln [ ψ ψ * ] t ) 1 ( ( 2 m c ) 2 ln [ ψ ψ * ] q μ ln [ ψ ψ * ] q ν + ( 1 V q u m c 2 ) g μ ν )  

Charged Boson in Non-Euclidean Space–Time

The KGE in non-Euclidean space–time for electromagnetic charged boson
1 g D μ ( g D μ ψ ) A μ μ g g = ψ ( m 2 c 2 2 + λ | ψ | 2 )
leads to the hydrodynamic system of equations
1 g μ g ( g μ ν | ψ | 2 ( ν + e A ν ) ) = 0
g μ ν ( μ S + e A μ ) ( ν S + e A ν ) = m 2 c 2 m V q u
where
V q u = 2 m ( 1 | ψ | g μ g ( g μ ν ν | ψ | ) + λ | ψ | 2 ) ,
Moreover, the Lagrangean motion equations read
p n μ = L n q ˙ μ D t p n μ = L n q μ ,
where
L n = ( ± ) ( m c 2 g μ ν q ˙ ν q ˙ μ c 2 1 + 2 m 2 c 2 ( μ μ | ψ n | | ψ n | + λ | ψ n | 2 ) g μ ν e A ν q ˙ μ ) ,
and to the QIETD
T n μ ν = ( ± ) m | ψ n | 2 c 2 g μ ν q ˙ ν q ˙ μ c 2 1 V q u ( n ) m c 2 ( ( u μ + ( 1 V q u ( n ) m c 2 ) 1 e m c A μ ) u ν ( 1 + ( 1 V q u ( n ) m c 2 ) 1 e m c g α β A β u α ) δ μ ν )

4. Discussion and Conclusions

Even if the GE for quantum systems (54) is formally equal to the Einstein GE, it actually demonstrates some differences. If the Einstein GE refers to the gravitational description of classical masses with no information, either about the quantum mechanical properties of bodies or about the string properties of elementary particles building up the matter, the GE (54) introduces the quantum properties in the gravity of a mass distribution that leads to a cosmological term, that is not a simple constant, but rather, has the form of an energy–impulse tensor density that leads to the correct value of the cosmological constant on the galactic space for the description of the motion of galaxies [68].
Even if the gravity of a quantum mechanical system refers to non-quantized fields, it is not without interest, since it is able to define the gravitational coupling of classical fields and their commutation rules in the non Euclidean space–time for the derivation of the quantum gravity equation of quantum fields.
The GE (54) together with (53) allows also to derive the mass eigenvalues of a Schwarzschild black hole [45,68] (like the e-atom model that is currently of interest to the scientific community [38] (and references therein)).
Finally, in this work the hydrodynamic model for bosons described by a KGE with perturbative self-interacting field is derived for defining the coupling with the Einstein gravitational equation of quantum mechanical systems.
The hydrodynamic treatment has a classical-like structure where the quantum dynamics (quantization) is introduced by the quantum potential through the factor 1 V q u ( n ) m c 2 .
This can be clearly seen by the Lagrangean function L n = ( ± ) m c 2 γ 1 V q u ( n ) m c 2 = ( ± ) m c 2 g μ ν q ˙ ν q ˙ μ c 2 1 V q u ( n ) m c 2 that, in the classical limit for 0 (i.e., V q u ( n ) = 0 and 1 V q u ( n ) m c 2 = 1 ), leads to the known classical expression.
Given the classical-like hydrodynamic model of quantum mechanics, the energy–impulse tensor for the Einstein gravitational equation of quantum mechanical systems is straightforwardly derived. The availability of such a description for mesons facilitates the coupling of them to the GE.
The biunique correspondence between the standard quantum mechanics and the hydrodynamic representation [60,61,62,63,70,71,77] warrants that the quantum energy–impulse tensor density is independent by the used formalism.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Rovelli, C. Quantum Gravity; Cambridge University Press: Cambridge, UK, 2004. [Google Scholar]
  2. Bousso, R. The Holographic principle. Rev. Mod. Phys. 2002, 74, 825. [Google Scholar] [CrossRef]
  3. Ashtekar, A. Introduction to loop quantum gravity and cosmology. In Quantum gravity and quantum cosmology; Springer: Berlin, Germany, 2013; pp. 31–56. [Google Scholar]
  4. Rovelli, C.; Vidotto, F. Covariant Loop Quantum Gravity; Cambridge University Press: Cambridge, UK, 2014. [Google Scholar]
  5. Thiemann, T. Modern Canonical Quantum General Relativity; Cambridge University Press: Cambridge, UK, 2007. [Google Scholar]
  6. Giddings, S.B. Is string theory a theory of quantum gravity? Found. Phys. 2013, 43, 115–139. [Google Scholar] [CrossRef]
  7. Dewitt, B.S. Quantum theory of gravity. I. The canonical theory. Phys. Rev. 1967, 160, 1113. [Google Scholar] [CrossRef]
  8. Gupta, S.N. Quantization of Einstein’s gravitational field: General treatment. Proc. Phys. Soc. Sect. A 1952, 65, 608–619. [Google Scholar] [CrossRef]
  9. Gupta, S.N.; Radford, S.F. Quantum field-theoretical electromagnetic and gravitational two-particle potentials. Phys. Rev. D 1980, 21, 2213. [Google Scholar] [CrossRef]
  10. Faddeev, L.D.; Popov, V.N. Feynman diagrams for the Yang-Mills field. Phys. Lett. B 1967, 25, 29–30. [Google Scholar] [CrossRef]
  11. Mandelstam, S. Feynman rules for the gravitational field from the coordinate-independent field-theoretic formalism. Phys. Rev. 1968, 175, 1604. [Google Scholar] [CrossRef]
  12. Schwinger, J. Quantized gravitational field. Phys. Rev. 1963, 130, 1253. [Google Scholar] [CrossRef]
  13. Schwinger, J. Sources and gravitons. Phys. Rev. 1968, 173, 1264–1272. [Google Scholar] [CrossRef]
  14. Weinberg, S. Photons and gravitons in S-matrix theory: Derivation of charge conservation and equality of gravitational and inertial mass. Phys. Rev. B 1964, 135, B1049. [Google Scholar] [CrossRef]
  15. Iwasaki, Y. Quantum theory of gravitation vs. classical theory fourth-order potential. Prog. Theor. Phys. 1971, 46, 1587–1609. [Google Scholar] [CrossRef]
  16. T Hooft, G.; Veltman, M.J.G. One-loop divergencies in the theory of gravitation. Ann. Henri Poincaré 1974, 20, 69–94. [Google Scholar]
  17. Radkowski, A.F. Some aspects of the source description of gravitation. Ann. Phys. 1970, 56, 319–354. [Google Scholar] [CrossRef]
  18. Capper, D.M.; Leibbrandt, G.; Ramon Medrano, M. Calculation of the graviton self-energy using dimensional regularization. Phys. Rev. D 1973, 8, 4320. [Google Scholar] [CrossRef]
  19. Duff, M.J. Quantum corrections to the Schwarzschild solution. Phys. Rev. D 1974, 9, 1837. [Google Scholar] [CrossRef]
  20. Capper, D.M.; Duff, M.J.; Halpern, L. Photon corrections to the graviton propagator. Phys. Rev. D 1974, 10, 461. [Google Scholar] [CrossRef]
  21. Capper, D.M.; Duff, M.J. The one loop neutrino contribution to the graviton propagator. Nucl. Phys. B 1974, 82, 147–154. [Google Scholar] [CrossRef]
  22. Duff, M.J.; Liu, J.T. Complementarity of the Maldacena and Randall-Sundrum pictures. Phys. Rev. Lett. 2000, 85, 2052. [Google Scholar] [CrossRef] [PubMed]
  23. Odintsov, S.D.; Shapiro, I.L. General relativity as the low-energy limit in higher derivative quantum gravity. Class. Quantum Gravity 1992, 9, 873. [Google Scholar] [CrossRef]
  24. Elizalde, E.; Odintsov, S.D.; Shapiro, I.L. Asymptotic regimes in quantum gravity at large distances and running Newtonian and cosmological constants. Class. Quantum Gravity 1994, 11, 1607. [Google Scholar] [CrossRef]
  25. Elizalde, E.; Lousto, C.O.; Odintsov, S.D.; Romeo, A. GUT’s in curved spacetime: Running gravitational constants, Newtonian potential, and the quantum-corrected gravitational equations. Phys. Rev. D 1995, 52, 2202. [Google Scholar] [CrossRef]
  26. Dalvit, D.A.R.; Mazzitelli, F.D. Running coupling constants, Newtonian potential, and nonlocalities in the effective action. Phys. Rev. D 1994, 50, 1001. [Google Scholar] [CrossRef]
  27. Dalvit, D.A.R.; Mazzitelli, F.D. Geodesics, gravitons, and the gauge-fixing problem. Phys. Rev. D 1997, 56, 7779. [Google Scholar] [CrossRef]
  28. Satz, A.; Mazzitelli, F.D.; Alvarez, E. Vacuum polarization around stars: Nonlocal approximation. Phys. Rev. D 2005, 71, 064001. [Google Scholar] [CrossRef]
  29. Anderson, P.R.; Fabbri, A. Apparent universality of semiclassical gravity in the far field limit. Phys. Rev. D 2007, 75, 044015. [Google Scholar] [CrossRef]
  30. Stephens, C.R.; Hooft, G.t.; Whiting, B.F. Black hole evaporation without information loss. Class. Quantum Gravity 1994, 11, 621. [Google Scholar] [CrossRef]
  31. Hawking, S.W. Particle creation by black holes. Commun. Math. Phys. 1975, 43, 199–220. [Google Scholar] [CrossRef]
  32. Parikh, M.K.; Wilczek, F. Hawking radiation as tunneling. Phys. Rev. Lett. 2000, 85, 5042–5045. [Google Scholar] [CrossRef] [PubMed]
  33. Parikh, M.K. A secret tunnel through the horizon. Gen. Relat. Gravitat. 2004, 36, 2419–2422. [Google Scholar] [CrossRef]
  34. Banerjee, R.; Majhi, B.R. Quantum tunneling beyond semiclassical approximation. J. High Energy Phys. 2008, 2008, 095. [Google Scholar] [CrossRef]
  35. Angheben, M.; Nadalini, M.; Vanzo, L.; Zerbini, S. Hawking radiation as tunneling for extremal and rotating black holes. J. High Energy Phys. 2005, 2005, 014. [Google Scholar] [CrossRef]
  36. Arzano, M.; Medved, A.J.M.; Vagenas, E.C. Hawking radiation as tunneling through the quantum horizon. J. High Energy Phys. 2005, 2005, 037. [Google Scholar] [CrossRef]
  37. Banerjee, R.; Majhi, B.R. Hawking black body spectrum from tunneling mechanism. Phys. Lett. B 2009, 675, 243–245. [Google Scholar] [CrossRef]
  38. Corda, C. Quasi-normal modes: The electrons of black holes as gravitational atoms? Implications for the black hole information puzzle. Adv. High Energy Phys. 2015, 2015, 867601. [Google Scholar] [CrossRef]
  39. Hajcek, P.; Kiefer, C. Singularity avoidance by collapsing shells in quantum gravity. Int. J. Mod. Phys. D 2001, 10, 775–779. [Google Scholar] [CrossRef]
  40. Bambi, C.; Malafarina, D.; Modesto, L. Non-singular quantum-inspired gravitational collapse. Phys. Rev. D 2013, 88, 044009. [Google Scholar] [CrossRef]
  41. Rovelli, C.; Vidotto, F. Planck stars. Int. J. Mod. Phys. D 2014, 23, 1442026. [Google Scholar] [CrossRef]
  42. Barrau, A.; Rovelli, C. Planck star phenomenology. Phys. Lett. B 2014, 739, 405–409. [Google Scholar] [CrossRef] [Green Version]
  43. Carr, B.J.; Mureika, J.; Nicolini, P. Sub-Planckian black holes and the Generalized Uncertainty Principle. J. High Energy Phys. 2015, 2015, 1–24. [Google Scholar] [CrossRef] [Green Version]
  44. T Hooft, G. The quantum black hole as a hydrogen atom: Microstates without strings attached. arXiv, 2016; arXiv:arXiv:1605.05119. [Google Scholar]
  45. Chiarelli, P. The quantum lowest limit to the black hole mass. Phys. Sci. Int. J. 2016, 9, 1–25. [Google Scholar] [CrossRef]
  46. Einstein, A. Zum Kosmologischen Problem der Allgemeinen Relativit Atstheorie; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 1931; Volume 142, pp. 235–237. [Google Scholar]
  47. Carroll, S.M.; Press, W.H.; Turner, E.L. The cosmological constant. Annu. Rev. Astron. Astrophys. 1992, 30, 499–542. [Google Scholar] [CrossRef]
  48. Zel’dovich, Y.B. The cosmological constant and the theory of elementary. Sov. Phys. Uspekhi 1968, 11, 381–393. [Google Scholar] [CrossRef]
  49. Gamov, G. My World Line; Viking Press: New York, NY, USA, 1970; p. 44. [Google Scholar]
  50. Hubble, E.P. A Relation between Distance and Radial Velocity among Extra-Galactic Nebulae. Proc. Natl. Acad. Sci. USA 1929, 15, 168–173. [Google Scholar] [CrossRef] [PubMed]
  51. Cohn, J.D. Living with Lambda. Astrophys. Space Sci. 1998, 259, 213–234. [Google Scholar] [CrossRef]
  52. Sahni, V.; Starobinsky, A.A. The case for a positive cosmological Λ-Term. Int. J. Mod. Phys. D 2000, 9, 373–443. [Google Scholar] [CrossRef]
  53. Turner, M.S. Dark matter and dark energy in the universe. In The Third Stromlo Symposium: The Galactic Halo, Proceedings of the Third Stromlo Symposium, Canberra, ACT, Australia, 17–21 August 1998; Gibson, B.K., Axelrod, T.S., Putman, M.E., Eds.; ASP Conference Series; ASP: San Francisco, CA, USA, 1998; Volume 165, pp. 431–452. [Google Scholar]
  54. Weinberg, S. The cosmological constant problem. Rev. Mod. Phys. 1989, 61, 1–23. [Google Scholar] [CrossRef]
  55. Dewitt, B.S. Quantum theory of gravity. II. The manifestly covariant theory. Phys. Rev. 1967, 162, 1195. [Google Scholar] [CrossRef]
  56. Veltman, M. Gravitation, in Les Houches; Balian, R., Zinn-Justin, J., Eds.; Session XXVIII; North-Holland Publishing Company: Amsterdam, the Netherlands, 1976; p. 266. [Google Scholar]
  57. Goroff, M.H.; Sagnotti, A. The ultraviolet behavior of Einstein gravity. Nucl. Phys. B 1986, 266, 709–736. [Google Scholar] [CrossRef]
  58. Van de Ven, A.E. Two-loop quantum gravity. Nucl. Phys. B 1992, 378, 309–366. [Google Scholar] [CrossRef]
  59. Akhundov, A.; Shiekh, A. A Review of leading quantum gravitational corrections to Newtonian gravity. arXiv, 2006; arXiv:arXiv:gr-qc/0611091. [Google Scholar]
  60. Bialyniki-Birula, I.; Cieplak, M.; Kaminski, J. Theory of Quanta; Oxford University Press: New York, NY, USA, 1992; pp. 87–115. [Google Scholar]
  61. Bohm, D. A suggested interpretation of the quantum theory in terms of “hidden” variables. I. Phys Rev. 1952, 85, 166. [Google Scholar] [CrossRef]
  62. Madelung, E.Z. Quantentheorie in hydrodynamischer Form. Z. Physik A 1926, 40, 322–326. [Google Scholar] [CrossRef]
  63. Jánossy, L. Zum hydrodynamischen Modell der Quantenmechanik. Z. Phys. 1962, 169, 79–89. [Google Scholar] [CrossRef]
  64. Aharonov, Y.; Bohm, D. Significance of electromagnetic potentials in quantum theory. Phys. Rev. 1959, 115, 485–491. [Google Scholar] [CrossRef]
  65. Weiner, J.H. Statistical Mechanics of Elasticity; John Wiley & Sons: New York, NY, USA, 1983; pp. 315–317. [Google Scholar]
  66. Haas, F.; Eliasson, B.; Shukla, P.K. Relativistic Klein-Gordon-Maxwell multistream model for quantum plasmas. Phys. Rev. E 2012, 85, 056411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Chiarelli, P. The CPT-Ricci scalar curvature symmetry in quantum electro-gravity. Int. J. Sci. 2016, 5, 36–58. [Google Scholar] [CrossRef]
  68. Chiarelli, P. The Quantum-Gravity Equation Derived from the Minimum Action Principle, Submitted for Publication. Available online: https://www.researchgate.net/publication/318420781_The_quantum-gravity_equation_derived_from_the_minimum_action_principle (accessed on 1 September 2017).
  69. Guvenis, H. Hydrodynamische Formulierung der relativischen Quantenmechanik. Available online: http://gsjournal.net/Science-Journals/Essays/View/5241 (accessed on 9 January 2014).
  70. Chiarelli, P. Theoretical derivation of the cosmological constant in the framework of the hydrodynamic model of quantum gravity: The solution of the quantum vacuum catastrophe? Galaxies 2016, 4, 6. [Google Scholar] [CrossRef]
  71. Weiner, J.H.; Askar, A. Particle method for the numerical solution of the time-dependent schrödinger equation. J. Chem. Phys. 1971, 54, 3534–3541. [Google Scholar] [CrossRef]
  72. Koide, T.; Kodama, T. Stochastic variational method as quantization scheme: Field quantization of complex Klein-Gordon equation. Prog. Theor. Exp. Phys. 2015. [CrossRef]
  73. Chiarelli, P. The Planck law for particles with rest mass. Quantum Matter 2016, 5, 748–751. [Google Scholar] [CrossRef]
  74. Hiley, B.J.; Callaghan, R.E. The Clifford Algebra approach to Quantum Mechanics A: The Schrödinger and Pauli Particles. arXiv, 2010; arXiv:arXiv:1011.4031. [Google Scholar]
  75. Landau, L.D.; Lifsits, E.M. Course of Theoretical Physics; Butterworth-Heinemann: Oxford, UK, 1976; Volume 2, pp. 309–335. [Google Scholar]
  76. Le Bellac, M. Quantum and Statistical Field Theory; Oxford Science Publication: Oxford, UK, 1991; pp. 315–337. [Google Scholar]
  77. Tsekov, R. Bohmian Mechanics versus Madelung Quantum Hydrodynamics. arXiv, 2009; arXiv:arXiv:0904.0723. [Google Scholar]

Share and Cite

MDPI and ACS Style

Chiarelli, P. The Non-Euclidean Hydrodynamic Klein–Gordon Equation with Perturbative Self-Interacting Field. Technologies 2017, 5, 63. https://doi.org/10.3390/technologies5040063

AMA Style

Chiarelli P. The Non-Euclidean Hydrodynamic Klein–Gordon Equation with Perturbative Self-Interacting Field. Technologies. 2017; 5(4):63. https://doi.org/10.3390/technologies5040063

Chicago/Turabian Style

Chiarelli, Piero. 2017. "The Non-Euclidean Hydrodynamic Klein–Gordon Equation with Perturbative Self-Interacting Field" Technologies 5, no. 4: 63. https://doi.org/10.3390/technologies5040063

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop