Cosmological constraints on dark matter particle production rate
Abstract
Gravitational particle production has been investigated
by using Einstein’s gravitational field equations in the presence
of a cosmological constant. To study the mechanism of particle creation,
the Universe has been considered as a thermodynamics system and
non-equilibrium thermodynamics has been employed.
In order to estimate the cosmological parameters with observational
data, including SNe Ia, BAO, Planck 2015 and HST, we have chosen a
phenomenological approach for the rate of particle creation.
A non-zero particle production rate was obtained implying
that the possibility of the particle production is consistent
with recent cosmological observations.
In the confidence interval, the ratio of
was obtained to be .
Keywords: Gravitational particle production, Particle production rate,
non-equilibrium thermodaynamics, Cosmological constraints.
I Introduction
The particle creation mechanism in cosmology was first introduced by Schrödinger Schrödinger . He investigated the effects of particle production on the evolution of the Universe by using the microscopic description of the gravitational particle production in an expanding Universe. About three decades later, Parker and others Parker ; Birrell ; Mukhanov ; Grib argued this idea again based on the quantum field theory in curved space-time with the motivation to find new consequences of the quantum field theory of fundamental particles. Parker combined quantum mechanics with general relativity and concluded that the time variations of the gravitational field lead to the production of the particle. After that, Hawking investigated particle production by black holes and its compatibility with the laws of thermodynamics Hawking .
In most cosmological models, the perfect fluid is taken into account while the real fluids are dissipative. Therefore, the description of many cosmological phenomena necessitates non-equilibrium or irreversible thermodynamics. Eckart Eckart and Landau and Lifschitz Landau1958 pioneered the generalization of irreversible thermodynamics from Newtonian fluid to relativistic fluid and considered the first order deviation from equilibrium. However, the first-order theory suffers from stability and causality problems.
The second-order deviation from equilibrium thermodynamics is considered by Muller ; Israel1976 ; Pavon ; Hiscock . Altering the dissipative phenomena into the dynamical variables that have the causal evolution overcome the causality problem, and the evolution equation limits the propagation speed of dissipative perturbations.
On the other hand, Prigogine Prigogine investigated the open thermodynamic system in the context of cosmology and concluded that although the particle production has not been achieved from Einstein’s gravitational field equations, particle production mechanism is consistent with these equations. In 1992, Calvao Pavon ; Calvao offered the covariant formulation of the particle production mechanisms.
The rate of produced particles should be determined by the quantum field theory in curved space-time Birrell . The exact functional form of the particle production rate is still not available; therefore, cosmologist have adopted the phenomenological approach and fitted it with observational data Steigman ; Lima ; Lima2014 ; Ramos .
In this work, our aim is to constrain the dark matter particle production rate with the observational data. In Section II, we apply non-equilibrium thermodynamics on the homogeneous and anisotropic background and study the particle production and its corresponding entropy production. The cosmological constraints used to estimate the free parameters of the model are presented in Section III. Eventually, we present the numerical results in Section IV. Discussion and conclusions are presented in Section V.
II Non-equilibrium thermodaynamics and gravitational particle production
Let us consider the expanding Universe with the line element including directional scale factors A(t), B(t) and C(t) known as Bianchi type I (BI) Bianchi
(1) |
The Einstein gravitational field equations in the presence of the cosmological constant are as follows,
(2) | |||
(3) | |||
(4) | |||
(5) |
where is the energy-momentum tensor and dot means differentiation with respect to the cosmic time t. By introducing the time-dependent function known as the vacuum scale of the BI Universe
(6) |
one can write metric functions explicitly. Also by this vacuum scale, the generalized Hubble parameter can be obtained
(7) |
(8) |
where are the integration constant. As , there is an isotropic expansion in all directions Jacobs . After the equations (8) are integrated, the explicit expression of the metric functions is obtained as follows, Saha
(9) |
Here and are arbitrary constants that satisfy the following relations:
(10) |
Finally, a little calculation on the equations (2-5) gives the evolution equation of Saha
(11) |
Modified energy-momentum tensor for relativistic fluid including particle production is given by the following equation
(12) |
where is the four-vector of velocity such that , and are energy density and equilibrium pressure,repectively, and is the pressure associated with the created particle. In a closed thermodynamic system, particle number () is conserved. Laws of energy conservation () and particle numbers conservation () lead to the following equations
(13) | |||
(14) |
where , and . In Israel authors introduce the second-order non-equilibrium thermodaynamics with entropy flow vector
(15) |
where entropy per particle is denoted by , fluid temperature by T, relaxation time by and coefficient of bulk viscosity by . Gibbs equation for this system
(16) |
gives the following relation for variation of the entropy per particle
(17) |
From the conservation laws (13,14) and (17) it is easy to show that
(18) |
With the following choice for creation pressure
(19) |
the second law of thermodynamics is satisfied Subhajit
(20) |
In open thermodynamic system, the particle number is not preserved () Zeldovich ; Prigogine2 ; thus, equation (13) should be modified as follows,
(21) |
where is the particle production rate. In such a case, the Gibbs equation gives
(22) |
which means that to recover the energy conservation equation (14), the entropy per particle must be constant (). Gibbs equation with conservation laws (14,21) gives
(23) |
so under the adiabatic condition, the creation pressure is as follows
(24) |
Therefore, in the adiabatic process, the creation pressure is linearly related to the particle production rate . So under the adiabatic condition, dissipative fluid is equivalent to a perfect fluid with variable particle number. On the other hand, from equations (11, 14, 24) it can be deduced that
(25) |
which means that regardless of the amount of equation of state, leads to a late time de Sitter (=0).
In an open thermodynamic system, the entropy change () in addition to the entropy flow () involves the entropy production () Harko
(26) |
where . In the homogeneous Universe, the entropy flow change is zero and so the entropy change is only due to the entropy production
(27) |
After integration, we have
(28) |
where and are the present value of entropy and scale factor, respectively.

In the BI Universe dominated by pressureless matter, baryonic matter and the energy of the quantum vacuum, Friedmann equation (11) is as follows
(29) | |||
only particle production for dark matter is considered here. To go ahead it is necessary to specify the particle production rate . In fact, is determined by studying the irreversible particle production in quantum field theory in curved space-time. The nature of the particles produced affects the particle production rate since the nature of the dark matter particles is not yet known, a phenomenological choice for particle production rate seems to be a feasible solution. A general phenomenological choice for particle production rate is where is an arbitrary function of the scale factor , and is a non-negative parameter. Following Nunes Nunes , we work with the phenomenological ansatz as follows,
(30) |


III Cosmological Constraint
In order to place cosmological constraints on the six free parameters of the model (II), including and , we run the CosmoMC package Lewis2000 ; Lewis2002 , that uses Markov Chain Monte Carlo (MCMC) algorithm to calculate the likelihood of cosmological parameters by using SNe Ia, BAO, Planck 2015 and HST observations. The likelihood function is defined as , such that represents the difference between observation and theory. The total likelihood is obtained by multiplying the separate likelihoods of SNe Ia, CMB, BAO, and HST data; thus, . For more details about cosmological constraint see Karami2013 ; Karami2014 .
III.1 Type Ia Supernovae (SNe Ia)
Type Ia supernovae have the same absolute magnitude and therefore these standard candles are a powerful tool for exploring the history of the expansion of the Universe. For our purpose, we employ Joint Light-curve Analysis (JLA) dataset, which in total comprises 740 SNe Ia data points in the redshift range Betoule . 118 SNe Ia within the redshift range from a combination of various subsamples Hamuy ; Riess ; Jha ; Contreras ; Hicken1 ; Hicken2 , 374 SNe Ia from Solon Digital Sky Survey (SDSS) within the redshift range Holtzman , 239 SNe Ia from the Supernova Legacy Survey (SNLS) within the redshift range Guy , and 9 SNe Ia from Hubble Space Telescope within the redshift range Riess2007 comprise the JLA collection.
III.2 Baryon Acoustic Oscillations (BAO)
BAO’s standard ruler has provided an other tool to probe the expansion history of the Universe. Cosmological perturbations in baryon-photon primordial plasma generate pressure waves that affect anisotropies of the CMB and the large scale structures of matter. The observed peak in the large scale correlation function measured by the luminous red galaxies of Solon Digital Sky Survey (SDSS) at z=0.35 Eisenstein and z=0.278 Kazin reveals the baryon acoustic oscillations at Mpc as well as in the two-degree Field Galaxy Redshift Survey (2dFGRS) at z=0.2 Percival , six-degree Field Galaxy Redshift Survey (6dFGRS) at z=0.106 Beutler , z=0.44, z=0.60 and z=0.73 by WiggleZ team Blake , the SDSS Data Releases 7 main Galaxy sample at z=0.15 Ross , the Data Releases 10 and 11 Galaxy samples at z=0.57 Anderson .
III.3 Cosmic Microwave Background (CMB)
Acoustic peaks of the temperature power spectrum of the cosmic microwave background radiation provide useful information about the expansion history of the Universe. The physics of decoupling affects the amplitude of the acoustic peaks and the physics of between the present and the decoupling changes the locations of peaks. Further on to probe the entire expansion history up to the last scattering surface, we will include the CMB data from Planck 2015 Adam ; Ade i.e. a joint observation of lowl + TT temperature fluctuations angular power spectrum.
III.4 Hubble Space Telescope (HST)
Another independent constraint that can be applied to the estimation of the model parameters is the Hubble parameter observational data obtained based on different ages of the galaxies Jimenez . Because the Hubble constant is included in many cosmological and astrophysical calculations, NASA/ESA built the Hubble Space Telescope (HST) to measure precise , and one of the three major HST projects was designated measurement with an accuracy of 10%. Freedman and his colleagues have obtained a new high-accuracy calibration of the Hubble constant based on our analysis of the Spitzer data available to date, combined with data from the Hubble Key Project. There was found a value of with a systematic uncertainty of and a statistical uncertainty of Freedman2012 . Riess et al. determined the Hubble constant from optical and infrared observations of over 600 Cepheid variables by using the Wide Field Camera 3 on the Hubble Space Telescope HST. They reported Hubble constant value of including systematic errors Riess2011 . Efstathiou reanalyzed the Riess et al. Riess2011 Cepheid data using the revised geometric maser distance to NGC 4258 of Humphreys et al. Humphreys . He concluded that based on the NGC 4258 maser distance is , compatible within with the recent determination from Planck, also assuming that the H-band period-luminosity relation is independent of metallicity Efstathiou .
IV Numerical results
We have constrained the parameters of the model, using observational data, including SNe Ia, BAO, Planck 2015 and HST. In Table 1, the main results of the statistical analysis are summarized.
Parameter |
|
Planck TT+lowl | ||
---|---|---|---|---|
In the confidence interval, the best fit of the parameter is obtained for Planck 2015, for joint analysis SNe Ia + BAO + HST + Planck 2015, which corresponds to and , respectively.
This result not only confirms the possibility of particle production, but shows that the cosmic scenario involving particle production is consistent with observational data.
Contribution of anisotropy that entered in the evolution of the background with parameter is estimated for Planck 2015, and for joint analysis of SNe Ia + BAO + HST + Planck 2015.
The values of Hubble constant for Planck 2015, and for joint analysis is consistent with Planck 2015 result Ade . The best fit of cosmological parameters, , , and is obtained , , and , for Planck 2015, and , and for joint analysis of SNe Ia + BAO + HST + Planck 2015. This result is consistent with the values repotred by Planck 2015 Ade .
In Figure 4, we have plotted the 1D likelihoods and 2D contours for model parameters with Planck TT + lowl (red), and SNe Ia + BAO + HST + Planck TT + lowl (blue), where contours represent confidence intervals of 68% and 95%.

Figure 5 displays the effective equation of state with the best fit of model parameters along with the effective equation of state of the model. The present value of the effective equation of state for this scenario is obtained . As the figure shows, at late time, the effective equation of state for this model tends to -1, which corresponds to late time de Sitter. In our study, this cosmic scenario showed a deviation from in the recent past.

Figure 6 shows the deceleration parameter with the best fit of model parameters along with the effective equation of state of the model. The present value of deceleration parameter for this scenario is obtained . The transition from deceleration to acceleration era occurred in . Our analysis of this cosmological model revealed a deviation from the model in the recent past.

In Figure 7, we plotted the evolutionary behaviors of the density parameters of Cold Dark Matter, , and vacuum density, . The figure displays the fact that as the density parameter of CDM is decreased, the vacuum density is increased, during the history of the Universe.

V Coclusions
In this paper, we examined the particle creation from the perspective of non-equilibrium thermodynamics in Bianchi type I Universe and tried to explain the evolution of the Universe by choosing a phenomenological approach for the particle production rate.
To constrain the free parameters of the model, we use the CosmoMC package. Our analysis includes observational data such as supernovae type Ia from JLA, cosmic microwave background from Planck 2015, baryon acoustic oscillation from SDSS, 2dFGRS, 6dFGRS, and observational Hubble data from HST. The results are as follows:
-
In confidence interval the best fit of the parameter is obtained for Planck 2015, for Planck 2015 + SNe Ia + BAO + HST. This value of the implies that and , respectively. Thus, the possibility of particle production is approved as consistent with recent cosmological observations.
-
Contribution of anisotropy in this model entered with obtained for Planck 2015 data is and for joint analysis Planck 2015 + SNe Ia + BAO + HST data is .
-
In this cosmic scenario the best fit of the cosmological parameters, , , and is consistent with similar values in the model. Figure 4 displayed 1D likelihoods and 2D contours for model parameters with Planck TT + low l (red), and joint analysis SNe Ia + BAO + HST + Planck TT + low l (blue), in and confidence intervals.
-
Effective equation of state for Planck 2015 data is , and for joint analysis Planck 2015 + SNe Ia + BAO + HST data is . Figure 5 shows EoS changes it is clear that at late time, it behaves like the model.
-
The deceleration parameter q reveals a transition from an early matter-dominant () epoch to the de Sitter era () at late time, as expected. The present value of the accelerating epoch starts at transition redshift . The deceleration parameter obtained.
-
Evolutionary behaviors of the density parameters of CDM, , and vacuum density, plotted in Figure 7, show that as the is increased, is decreased, during the history of the Universe.
References
- (1) E. Schrödinger, Physica (Amsterdam) 6, 899 (1939).
- (2) L. Parker, Phys. Rev. Lett. 21, 562 (1968).
- (3) N. D. Birrell and P. C. W. Davies, Quantum fields in curved space, Cambridge Univ. Press, Cambridge, England, 1982.
- (4) V. F. Mukhanov and S. Winitzki, Introduction to Quantum Effects in Gravity, Cambridge Univ. Press, Cambridge, England, 2007.
- (5) A. A. Grib, S. G. Mamayev, and V. M. Mostepanenko, Vacuum quantum effects in strong fields, Friedmann Laboratory Publishing, St. Petersburg, 1994.
- (6) S. W. Hawking, Commun. math. Phys. 43, 199 (1975).
- (7) C. Eckart, Phys. Rev. 58, 919 (1940).
- (8) L. D. Landau and E. M. Lifshitz, Fluid Mechanics, Addison-wesley, Reading, MA, 1994.
- (9) I. Muller, A. phys. 198, 329 (1967).
- (10) W. Israel, Ann. Phys. (N.Y) 100, 310 (1976).
- (11) D. Pavon, D. Jou, and J. Casas-Vazquez, Ann. Inst. Henri Pioncare A36, 79 (1982).
- (12) W. A. Hiscock, L. LindblomJou, and J. Casas-Vazquez, Phys. Rev. D 35, 3723 (1987).
- (13) I. Prigogine, J. Geheniau, E. Gunzig, and P. Nardone, Gen. Relativ. Gravit. 21, 767 (1989).
- (14) M. O. Calvao, J. A. S. Lima, and I. Waga, Phys. lett. A 162, 223 (1992).
- (15) G. Steigman, R. C. Santos, and J. A. S. Lima, J. Cosmol. Astropart. Phys 06, 033 (2009).
- (16) J. A. S. Lima, J. F. Jesus, and F. A. Oliveria, J. Cosmol. Astropart. Phys 11, 027 (2010).
- (17) J. A. S. Lima, L. L. Graef, D. Pavon, and S. Basilakos, J. Cosmol. Astropart. Phys 10, 042 (2014).
- (18) R. O. Ramos, M. V. Santos, and I. Waga, Phys. Rev. D 89, 083524 (2014).
- (19) L. Bianchi, Lezioni sulla teoria dei gruppi continui finiti di trasformazioni, Pisa, 1918.
- (20) K. C. Jacobs, Astrophysical Jurnal 153, 661 (1968).
- (21) B. Saha, Phys. Rev. D 64, 123501 (2001).
- (22) W. Israel and J. M. Stewart, Ann. Phys. (N.Y) 118, 341 (1979).
- (23) S. Saha and S. Chakraborty, Gen. Relativ. Gravit. 47, 127 (2015).
- (24) Y. B. Zeldovich, JETP Lett. 12, 307 (1970).
- (25) I. Prigogine, J. Geheniau, E. Gunzing, and P. Nardone, proc. Natl Acad. Sci 85, 7248 (1998).
- (26) T. Harko and F. S. N. Lobo, Phys. Rev. D 87, 044018 (2013).
- (27) R. C. Nunes and D. Pavon, Phys. Rev. D 91, 063526 (2015).
- (28) A. Lewis, A. Challinor, and L. A, Astrophys. J 538, 473 (2000).
- (29) A. Lewis and S. Bridle, Phys. Rev. D 66, 103511 (2002).
- (30) K. Karami, S. Asadzadeh, A. Abdolmaleki, and Z. Safari, Phys. Rev. D 88, 084034 (2013).
- (31) S. Asadzadeh, Z. Safari, K. Karami, and A. Abdolmaleki, Int. J. Theor. Phys. 53, 1248 (2014).
- (32) M. Betoule et al., Astron. Astrophys. 568, A22 (2014).
- (33) M. Hamuy et al., Astron. J. 112, 2408 (1996).
- (34) A. G. Riess et al., Astron. J. 117, 707 (1999).
- (35) S. Jha et al., Astron. J. 131, 527 (2006).
- (36) M. Contreras et al., Astron. J. 139, 519 (2010).
- (37) M. Hicken et al., Astrophys. J. 700, 1097 (2009).
- (38) M. Hicken et al., Astrophys. J. 700, 331 (2009).
- (39) J. Holtzman et al., Astron. J. 136, 2306 (2008).
- (40) J. Guy et al., Astron. Astrophys. 253, 7 (2010).
- (41) A. G. Riess et al., Astrphys. J. 659, 98 (2007).
- (42) D. J. Eisenstein et al., Astrphys. J. 633, 560 (2005).
- (43) E. A. Kazin et al., Astrphys. J. 710, 1444 (2010).
- (44) W. J. Percival et al., Mon. Not. R. Astron. Soc. 401, 2148 (2010).
- (45) F. Beutler et al., Mon. Not. R. Astron. Soc. 416, 3017 (2011).
- (46) A. Blake et al., Mon. Not. R. Astron. Soc. 418, 1707 (2011).
- (47) A. J. Ross et al., Mon. Not. R. Astron. Soc. 449, 835 (2015).
- (48) L. Anderson et al., Mon. Not. R. Astron. Soc. 441, 24 (2014).
- (49) R. Adam et al., Astron. Astrophys. 594, A1 (2016).
- (50) P. A. R. Ade et al., Astron. Astrophys. 594, A13 (2016).
- (51) R. Jimenez and A. Loeb, Astrphys. J. 573, 37 (2002).
- (52) W. L. Freedman et al., Astrphys. J. 758, 24 (2012).
- (53) A. G. Riess et al., Astrphys. J. 730, 119 (2011).
- (54) E. M. L. Humphreys et al., Astrphys. J. 755, 13 (2013).
- (55) G. Efstathiou, Mon. Not. R. Astron. Soc. 440, 1138 (2014).