This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Second order hydrodynamics based on effective kinetic theory and electromagnetic signals from QGP

Lakshmi J. Naik [email protected]    V. Sreekanth [email protected] Department of Sciences, Amrita School of Physical Sciences, Coimbatore, Amrita Vishwa Vidyapeetham, India
Abstract

We study the thermal dilepton and photon production from relativistic heavy ion collisions in presence of viscosities by employing the recently developed second order dissipative hydrodynamic formulation estimated within a quasiparticle description of thermal QCD (Quantum Chromo-Dynamics) medium. The sensitivity of shear and bulk viscous pressures to the temperature dependence of relaxation time is analyzed within one dimensional boost invariant expansion of quark gluon plasma (QGP). The dissipative corrections to the phase-space distribution functions upto first order in gradients are obtained from the Chapman-Enskog like iterative solution of effective Boltzmann equation in the relaxation time approximation. Thermal dilepton and photon production rates for QGP are calculated by employing this viscous modified distribution function. Yields of these particles are quantified for the longitudinal expansion of QGP with different temperature dependent relaxation times. Our analysis employing this second order hydrodynamic model indicates that the spectra of dileptons and photons gets enhanced by both bulk and shear viscosities and is well behaved. Also, these particle yields are found to be sensitive to relaxation time. Further, we do a comparison of these particle spectra with a standard hydrodynamic formulation.

I Introduction

The experiments at Relativistic Heavy Ion Collider (RHIC) and at Large Hadron Collider (LHC) suggest the existence of strongly coupled quark-gluon matter at extreme high temperature and density Adams et al. (2005); Adcox et al. (2005); Arsene et al. (2005); Back et al. (2005). These experiments provide opportunity to inspect the features of the hot nuclear matter, QGP which is believed to have existed in the primordial universe. Theoretical and experimental investigations of properties of the hot and dense QGP is being pursued vigorously by the community Jaiswal et al. (2021). Analysis of the experimental data from the heavy-ion collision experiments imply that the QGP has a near perfect fluid nature with extreme low value of shear viscosity to entropy ratio, η/s=1/4π\eta/s=1/4\pi Kovtun et al. (2005). This surprisingly low value of viscosity has evoked interest in the application of relativistic dissipative hydrodynamics to heavy ion collisions Romatschke and Romatschke (2019).

The evolution of viscous QGP has to be modelled using higher order relativistic viscous hydrodynamical theories, since first order Navier-Stokes theory exhibits acausal behaviour and numerical instabilities Hiscock and Lindblom (1983, 1985); Geroch and Lindblom (1990). Recently, several investigations are being pursued towards the development of causal first order hydrodynamic theory Bemfica et al. (2018); Hoult and Kovtun (2020); Biswas et al. (2022). When it comes to second order theories, there is no unique prescription and there exist several successful formalisms. The earliest attempts in this direction were by Muller, Israel and Stewart MÃŒller (1967); Israel and Stewart (1979). Development of second order causal theories is an active field of research and several new formalisms have been proposed and studied in the context of heavy ion collisions Jaiswal and Roy (2016); Romatschke and Romatschke (2019). Second order viscous hydrodynamics developed in Ref. Bhadury et al. (2021) is a recently proposed formalism within the effective fugacity quasiparticle model (EQPM) for the hot QCD medium Chandra and Ravishankar (2009, 2011). The EQPM prescription incorporates the thermal QCD medium interaction effects into a system of quasipartons by considering thermal modifications to the phase-space distribution functions in terms of effective quark and gluon fugacities. The dissipative hydrodynamic evolution equations have been determined by employing the effective covariant kinetic theory developed for EQPM Mitra and Chandra (2018). The non-equilibrium corrections to the distribution functions have been estimated in this hydrodynamic framework by the Chapman-Enskog (CE) expansion in relaxation time approximation (RTA) Jaiswal (2013a, b). Moreover, a quasiparticle model such as EQPM distinguishes the quark and gluon sector in the hot QCD medium and can be used to study thermal dilepton and photon emission from QGP Chandra and Sreekanth (2015).

Properties of the QCD matter created in high energy heavy ion collisions can be studied by analysing the different signals emitted, such as thermal dileptons and photons. As these particles interact only electromagnetically, they can easily decouple from the strong nuclear matter and reach the detectors without further interaction with other particles. Effect of viscosities has consequences on the thermal dilepton and photon spectra from heavy ion collisions. The role of shear viscosity on thermal particle production using causal dissipative hydrodynamics has been investigated in Refs. Dusling and Lin (2008); Dusling (2010); Bhatt and Sreekanth (2010); Bhatt et al. (2012); Chaudhuri and Sinha (2011); Vujanovic et al. (2014, 2018). The influence of bulk viscosity on the expansion of QGP and signals emanating from it has been studied in Refs. Torrieri et al. (2008); Fries et al. (2008); Rajagopal and Tripuraneni (2010); Bhatt et al. (2010, 2012); Paquet et al. (2016). In certain situations, the inclusion of dissipation into the heavy ion collision scenario may induce a phenomenon called cavitation which causes the hydrodynamic description to be invalid before the freeze-out and thereby affecting the signals Rajagopal and Tripuraneni (2010); Bhatt et al. (2011). Recently, thermal particle production has been investigated by employing the concept of hydrodynamic attractors for temperature evolution Coquet et al. (2021); Naik et al. (2021a, b). The EQPM prescription has been employed to study several observables from heavy ion collisions. Impact of chromo-Weibel instability on thermal dilepton emission has been analysed by the authors of Ref. Chandra and Sreekanth (2017) within EQPM. The effect due to collisional contributions of thermal QCD medium has also been investigated on thermal dileptons Naik et al. (2022) and heavy quark transport Prakash et al. (2021), using EQPM. However, in these studies employing the EQPM, ideal hydrodynamics was used to model the evolution of the system.

In the present analysis, we proceed to implement the causal second order hydrodynamic framework within EQPM to study the evolution of QGP and thermal particle production from heavy ion collisions. We analyze the evolution for different temperature dependent relaxation times. The thermal particle emission rates have to be calculated by proper modelling of the momentum distribution functions incorporating the viscous effects. The non-equilibrium part of distribution function is generally determined by the 14-moment Grad’s method or the CE expansion, out of these, the CE type expansion of non-equilibrium distribution function is shown to be well behaved even up to second order in gradients Bhalerao et al. (2014). In this work, we intend to calculate thermal dilepton and photon production rates using the viscous modified quasiparticle thermal distribution functions determined by the CE method in RTA. The particle emission spectra is dependent on the temperature profile of QGP and is found by modelling the expansion of QGP with relativistic dissipative hydrodynamics under one-dimensional (11-D) boost invariant Björken expansion.

The paper is structured as follows. In section II, we review the EQPM description and the formalism used to derive the second order viscous hydrodynamics. The causal hydrodynamic evolution equations of the expanding QGP medium are prescribed within the 1-D boost invariant Björken flow. Then, we analyze the evolution of shear stress and bulk viscous pressure for different relaxation times. Section III is devoted to the calculation of thermal dilepton and photon production rates in the presence of viscous modified momentum distribution function. In section IV, thermal particle spectra is computed in 1-D Björken expansion and we present the results of our analysis in section V. Section VI details conclusions and future outlook.

Notations and conventions: We are working with natural units i.e.,c==kB=1i.e.,\,c=\hbar=k_{B}=1. The Minkowski metric is taken as ημν=diag(1,1,1,1)\eta^{\mu\nu}=\textrm{diag}(1,-1,-1,-1). The fluid four-velocity is denoted as uμu^{\mu}, which is normalized as uμuμ=1u^{\mu}u_{\mu}=1 and in the local rest frame (LRF) of the fluid, uμ=(1,0,0,0)u^{\mu}=(1,0,0,0). The quantity Δμν=ημνuμuν\Delta^{\mu\nu}=\eta^{\mu\nu}-u^{\mu}u^{\nu} represents the projection operator orthogonal to uμu^{\mu} and μ=Δμνν\nabla^{\mu}=\Delta^{\mu\nu}\partial_{\nu}. Also, Δαβμν12(ΔαμΔβν+ΔβμΔαν)13ΔμνΔαβ\Delta_{\alpha\beta}^{\mu\nu}\equiv\frac{1}{2}(\Delta_{\alpha}^{\mu}\Delta_{\beta}^{\nu}+\Delta_{\beta}^{\mu}\Delta_{\alpha}^{\nu})-\frac{1}{3}\Delta^{\mu\nu}\Delta_{\alpha\beta} is the traceless symmetric projection operator orthogonal to uμu^{\mu}.

II Second order dissipative hydrodynamics based on quasiparticle model

In this section, we present the formalism to estimate the second order relativistic viscous hydrodynamic equations within the EQPM description of QCD medium as derived in Ref. Bhadury et al. (2021). The EQPM maps the thermal QCD medium effects through temperature dependent effective fugacity parameters of constituent noninteracting quasiparticles. Within this model, the equilibrium momentum distribution functions of the quasiparticles are given by Chandra and Ravishankar (2009),

fk0=zkexp[β(uμpkμ)]1±zkexp[β(uμpkμ)].f_{k}^{0}=\frac{z_{k}\exp[-\beta(u_{\mu}p_{k}^{\mu})]}{1\pm z_{k}\exp[-\beta(u_{\mu}p_{k}^{\mu})]}. (1)

Here, k(q,g)k\equiv(q,g) represent the quarks and gluons respectively and zq,zgz_{q},z_{g} denote the effective quark, gluon fugacities which capture the QCD medium interactions. The model divides the hot QCD medium into two sectors : (i) the effective gluonic sector, which describes the contribution of gluonic action to the pressure as well as the contribution due to the internal fermion lines, (ii) the matter sector, which represents the interactions between quarks and antiquarks, along with their interactions with gluons. The form of zgz_{g} and zqz_{q} is determined by equating the expression for pressure obtained within EQPM with the lattice data (see Ref. Chandra and Ravishankar (2011) for detailed discussion). Here, the (2+12+1) flavor lattice QCD equation of state Cheng et al. (2008); Borsanyi et al. (2014) is considered. In Eq. (1), pkμ=(Ek,pk)p_{k}^{\mu}=(E_{k},\vec{p}_{k}) represent the particle four-momenta of quarks and gluons and β=1/T\beta=1/T is the inverse of temperature. The quasiparticle momenta (p~kμ\tilde{p}_{k}^{\mu}) and bare particle momenta (pkμp_{k}^{\mu}) are related through the dispersion relation:

p~g,qμ=pg,qμ+δωg,quμ;δωg,q=T2Tln(zg,q),\tilde{p}_{g,q}^{\mu}=p_{g,q}^{\mu}+\delta\omega_{g,q}u^{\mu};\,\,\,\,\,\,\,\,\,\,\,\delta\omega_{g,q}=T^{2}\partial_{T}\ln(z_{g,q}), (2)

where δωg,q\delta\omega_{g,q} denotes the modified part of the dispersion relation. From Eq. (2), the single particle energy of the quasiparticles is given by

p~g,q0ωg,q=Eg,q+δωg,q.\tilde{p}_{g,q}^{0}\equiv\omega_{g,q}=E_{g,q}+\delta\omega_{g,q}. (3)

In order to obtain the viscous hydrodynamic equations, one need to quantify the dissipative corrections of the system. The non-equilibrium corrections to the phase space distribution functions are estimated by considering the effective Boltzmann equation within EQPM. The form of relativistic transport equation in RTA for the collision term is obtained as Mitra and Chandra (2018)

p~kμμfk0(x,p~k)+Fkμμ(p)fk0=δfkτRωk,\tilde{p}_{k}^{\mu}\partial_{\mu}f_{k}^{0}(x,\tilde{p}_{k})+F_{k}^{\mu}\partial_{\mu}^{(p)}f_{k}^{0}=-\frac{\delta f_{k}}{\tau_{R}}\omega_{k}, (4)

where τR\tau_{R} is the relaxation time and δfk\delta f_{k} denotes the non-equilibrium part of the distribution function. The force term is defined from energy-momentum and particle flow conservation and is given by Fkμ=ν(δωkuνuμ)F_{k}^{\mu}=-\partial_{\nu}(\delta\omega_{k}u^{\nu}u^{\mu}) Mitra and Chandra (2018).

Using the Landau definition : uνTμν=ϵuμu_{\nu}T^{\mu\nu}=\epsilon u^{\mu} Landau and Lifshitz (1987), the energy-momentum tensor can be decomposed in terms of hydrodynamic degrees of freedom as

Tμν=ϵuμuνPΔμν+τμν,T^{\mu\nu}=\epsilon u^{\mu}u^{\nu}-P\Delta^{\mu\nu}+\tau^{\mu\nu}, (5)

where ϵ\epsilon and PP denote the equilibrium energy density and pressure of the system respectively. τμν=πμνΠΔμν\tau^{\mu\nu}=\pi^{\mu\nu}-\Pi\Delta^{\mu\nu} is the dissipative current, with πμν\pi^{\mu\nu} being the traceless part. The quantities πμν\pi^{\mu\nu} and Π\Pi represent the shear stress tensor and bulk viscous pressure respectively. The evolution equations for ϵ\epsilon and uμu^{\mu} can be obtained by projecting the energy-momentum conservation equation μTμν=0\partial_{\mu}T^{\mu\nu}=0 along and orthogonal to uμu^{\mu} and are given by

ϵ˙+(ϵ+P+Π)Θπμνσμν\displaystyle\dot{\epsilon}+(\epsilon+P+\Pi)\Theta-\pi^{\mu\nu}\sigma_{\mu\nu} =\displaystyle= 0,\displaystyle 0, (6)
(ϵ+P+Π)u˙αα(P+Π)+Δναμπμν\displaystyle(\epsilon+P+\Pi)\dot{u}^{\alpha}-\nabla^{\alpha}(P+\Pi)+\Delta_{\nu}^{\alpha}\partial_{\mu}\pi^{\mu\nu} =\displaystyle= 0.\displaystyle 0. (7)

Here, Θμuμ\Theta\equiv\partial^{\mu}u_{\mu} is the four-divergence of fluid velocity, σμνΔαβμναuβ\sigma^{\mu\nu}\equiv\Delta_{\alpha\beta}^{\mu\nu}\nabla^{\alpha}u^{\beta} represents the shear stress tensor and A˙uμμA\dot{A}\equiv u^{\mu}\partial_{\mu}A is the co-moving derivative of AA. The derivatives of β\beta can be calculated from the above equations

β˙\displaystyle\dot{\beta} =\displaystyle= βcs2(Θ+ΠΘπμνσμνϵ+P),\displaystyle\beta c_{s}^{2}\left(\Theta+\frac{\Pi\Theta-\pi^{\mu\nu}\sigma_{\mu\nu}}{\epsilon+P}\right), (8)
αβ\displaystyle\nabla^{\alpha}\beta =\displaystyle= β(u˙α+Πu˙ααΠ+Δναμπμνϵ+P),\displaystyle-\beta\left(\dot{u}^{\alpha}+\frac{\Pi\dot{u}^{\alpha}-\nabla^{\alpha}\Pi+\Delta_{\nu}^{\alpha}\partial_{\mu}\pi^{\mu\nu}}{\epsilon+P}\right), (9)

where cs2=dP/dϵc_{s}^{2}=dP/d\epsilon denotes the speed of sound squared.

Now, the form of energy-momentum tensor in terms of quasiparticle four-momenta can be written as

Tμν(x)\displaystyle T^{\mu\nu}(x) =\displaystyle= kgk𝑑P~kp~kμp~kνfk(x,p~k)\displaystyle\sum_{k}g_{k}\int d\tilde{P}_{k}\tilde{p}_{k}^{\mu}\tilde{p}_{k}^{\nu}f_{k}(x,\tilde{p}_{k}) (10)
+kgkδωk𝑑P~k<p~kμp~kν>Ekfk(x,p~k),\displaystyle+\sum_{k}g_{k}\delta\omega_{k}\int d\tilde{P}_{k}\frac{<\tilde{p}_{k}^{\mu}\tilde{p}_{k}^{\nu}>}{E_{k}}f_{k}(x,\tilde{p}_{k}),

with gkg_{k} being the degeneracy factor of quasiparticles. Here, <p~kμp~kν>12(ΔαμΔβν+ΔβμΔαν)p~kαp~kβ<\tilde{p}_{k}^{\mu}\tilde{p}_{k}^{\nu}>\equiv\frac{1}{2}(\Delta_{\alpha}^{\mu}\Delta_{\beta}^{\nu}+\Delta_{\beta}^{\mu}\Delta_{\alpha}^{\nu})\tilde{p}_{k}^{\alpha}\tilde{p}_{k}^{\beta} and dP~kd3|pk|(2π)3ωkd\tilde{P}_{k}\equiv\frac{d^{3}|\vec{p}_{k}|}{(2\pi)^{3}\omega_{k}} represents the phase-space factor. Using Eqs. (5) and (10), the expressions for shear stress tensor and bulk viscous pressure are obtained as Mitra and Chandra (2018)

πμν\displaystyle\pi^{\mu\nu} =\displaystyle= kgkΔαβμν𝑑P~kp~kαp~kβδfk\displaystyle\sum_{k}g_{k}\Delta_{\alpha\beta}^{\mu\nu}\int d\tilde{P}_{k}\tilde{p}_{k}^{\alpha}\tilde{p}_{k}^{\beta}\delta f_{k} (11)
+kgkδωkΔαβμν𝑑P~kp~kαp~kβδfkEk,\displaystyle+\sum_{k}g_{k}\delta\omega_{k}\Delta_{\alpha\beta}^{\mu\nu}\int d\tilde{P}_{k}\tilde{p}_{k}^{\alpha}\tilde{p}_{k}^{\beta}\frac{\delta f_{k}}{E_{k}},
Π\displaystyle\Pi =\displaystyle= 13kgkΔαβ𝑑P~kp~kαp~kβδfk\displaystyle-\frac{1}{3}\sum_{k}g_{k}\Delta_{\alpha\beta}\int d\tilde{P}_{k}\tilde{p}_{k}^{\alpha}\tilde{p}_{k}^{\beta}\delta f_{k} (12)
13kgkδωkΔαβ𝑑P~kp~kαp~kβδfkEk.\displaystyle-\frac{1}{3}\sum_{k}g_{k}\delta\omega_{k}\Delta_{\alpha\beta}\int d\tilde{P}_{k}\tilde{p}_{k}^{\alpha}\tilde{p}_{k}^{\beta}\frac{\delta f_{k}}{E_{k}}.

The transport coefficients of the system can be determined once we know the form of δfk\delta f_{k}. In Ref. Bhadury et al. (2021), viscous correction is calculated by employing the Chapman-Enskog like iterative solution of the Boltzmann equation in RTA. Within the effective kinetic theory considered, the non-equilibrium corrections to the quark (antiquark) distribution function up to first order in gradients is then obtained as

δfq\displaystyle\delta f_{q} =\displaystyle= τR[p~qμμβ+βp~qμp~qνup~qμuνβΘ(δωq)\displaystyle\tau_{R}\bigg{[}\tilde{p}_{q}^{\mu}\partial_{\mu}\beta+\frac{\beta\,\tilde{p}_{q}^{\mu}\,\tilde{p}_{q}^{\nu}}{u\!\cdot\!\tilde{p}_{q}}\partial_{\mu}u_{\nu}-\beta\Theta(\delta\omega_{q}) (13)
ββ˙((δωq)β)]fq0f¯q0,\displaystyle-\beta\dot{\beta}\left(\frac{\partial(\delta\omega_{q})}{\partial\beta}\right)\bigg{]}f_{q}^{0}\bar{f}_{q}^{0},

where f¯q0=1fq0\bar{f}_{q}^{0}=1-f_{q}^{0}. By employing the above equation in Eqs. (11) and (12), and considering τR\tau_{R} to be independent of momenta, the following relations are obtained:

πμν=2τRβπσμν,Π=τRβΠΘ.\displaystyle\pi^{\mu\nu}=2\tau_{R}\beta_{\pi}\sigma^{\mu\nu},\,\,\,\,\,\,\,\,\,\,\Pi=-\tau_{R}\beta_{\Pi}\Theta. (14)

We note that, as a result of RTA, there will be only a single time-scale for both shear and bulk relaxation times. First order transport coefficients within the effective covariant kinetic theory can be determined by comparing the above equations with that of the Navier-Stokes equation

η=τRβπ,ζ=τRβΠ.\eta=\tau_{R}\beta_{\pi},\,\,\,\,\,\,\,\,\,\zeta=\tau_{R}\beta_{\Pi}. (15)

Here, η\eta and ζ\zeta are the coefficients of shear and bulk viscosities respectively. The quantities βπ,βΠ\beta_{\pi},\beta_{\Pi} appearing in the above equations are determined in terms of thermodynamic integrals and are given as

βπ\displaystyle\beta_{\pi} =\displaystyle= βk[J~k42(1)+δωkL~k42(1)],\displaystyle\beta\sum_{k}\bigg{[}\tilde{J}^{(1)}_{k~{}42}+\delta\omega_{k}\tilde{L}^{(1)}_{k~{}42}\bigg{]}, (16)
βΠ\displaystyle\beta_{\Pi} =\displaystyle= βk[cs2(J~k31(0)+δωkL~k31(0)\displaystyle\beta\sum_{k}\bigg{[}c_{s}^{2}\bigg{(}\tilde{J}_{k~{}31}^{(0)}+\delta\omega_{k}\tilde{L}_{k~{}31}^{(0)}
(δωk)β(J~k21(0)+δωkL~k21(0)))\displaystyle-\frac{\partial(\delta\omega_{k})}{\partial\beta}\left(\tilde{J}_{k~{}21}^{(0)}+\delta\omega_{k}\tilde{L}_{k~{}21}^{(0)}\right)\bigg{)}
+53(J~k42(1)+δωkL~k42(1))δωk(J~k21(0)+δωkL~k21(0))].\displaystyle+\frac{5}{3}\bigg{(}\tilde{J}_{k~{}42}^{(1)}+\delta\omega_{k}\tilde{L}_{k~{}42}^{(1)}\bigg{)}-\delta\omega_{k}\left(\tilde{J}_{k~{}21}^{(0)}+\delta\omega_{k}\tilde{L}_{k~{}21}^{(0)}\right)\bigg{]}.

The form of thermodynamic integrals, J~knm(r)\tilde{J}^{(r)}_{k~{}nm} and L~knm(r)\tilde{L}^{(r)}_{k~{}nm} appearing in the above equations are given in Appendix A. Following the methodology of Ref. Bhadury et al. (2021), the evolution equations for shear stress tensor and bulk viscous pressure are obtained as

π˙μν+πμντR\displaystyle\dot{\pi}^{\langle\mu\nu\rangle}+\frac{\pi^{\mu\nu}}{\tau_{R}} =\displaystyle= 2βπσμν+2πϕμωνϕδπππμνθ\displaystyle 2\beta_{\pi}\sigma^{\mu\nu}+2\pi_{\phi}^{\langle\mu}\omega^{\nu\rangle\phi}-\delta_{\pi\pi}\pi^{\mu\nu}\theta (18)
τπππϕμσνϕ+λπΠΠσμν,\displaystyle-\tau_{\pi\pi}\pi_{\phi}^{\langle\mu}\sigma^{\nu\rangle\phi}+\lambda_{\pi\Pi}\Pi\sigma^{\mu\nu},
Π˙+ΠτR\displaystyle\dot{\Pi}+\frac{\Pi}{\tau_{R}} =\displaystyle= βΠθδΠΠΠθ+λΠππμνσμν.\displaystyle-\beta_{\Pi}\theta-\delta_{\Pi\Pi}\Pi\theta+\lambda_{\Pi\pi}\pi^{\mu\nu}\sigma_{\mu\nu}. (19)

Here, ωμν=12(μuννuμ)\omega^{\mu\nu}=\frac{1}{2}(\nabla^{\mu}u^{\nu}-\nabla^{\nu}u^{\mu}) denotes the vorticity tensor. The second order transport coefficients in Eqs. (18) and (19) are obtained in terms of different thermodynamic integrals and are shown in Appendix A.

II.1 Viscous evolution for different relaxation times

We solve the second order viscous evolution equations, Eqs. (18) and (19), by choosing different temperature dependent forms for the relaxation time. We model the expansion using the 1-D Björken flow Bjorken (1983), which considers the QGP medium as a transversely homogenous and longitudinally boost-invariant system. It is now convenient to parameterize the coordinates as t=τcoshηst=\tau\cosh\eta_{s} and z=τsinhηsz=\tau\sinh\eta_{s}, where τ=t2z2\tau=\sqrt{t^{2}-z^{2}} is the proper time and ηs=12lnt+ztz\eta_{s}=\frac{1}{2}\ln\frac{t+z}{t-z} is the space-time rapidity of the system. Fluid four-velocity is expressed using the ansatz, uμ=(coshηs,0,0,sinhηs)u^{\mu}=(\cosh\eta_{s},0,0,\sinh\eta_{s}). Now, within this model, the hydrodynamic equations given by Eqs. (18) and (19) reduce to the following coupled non-linear differential equations in τ\tau Bhadury et al. (2021):

dϵdτ\displaystyle\frac{d\epsilon}{d\tau} =\displaystyle= 1τ(ϵ+P+Ππ),\displaystyle-\frac{1}{\tau}\left(\epsilon+P+\Pi-\pi\right), (20)
dπdτ+πτπ\displaystyle\frac{d\pi}{d\tau}+\frac{\pi}{\tau_{\pi}} =\displaystyle= 43βπτ(13τππ+δππ)πτ+23λπΠΠτ,\displaystyle\frac{4}{3}\frac{\beta_{\pi}}{\tau}-\left(\frac{1}{3}\tau_{\pi\pi}+\delta_{\pi\pi}\right)\frac{\pi}{\tau}+\frac{2}{3}\lambda_{\pi\Pi}\frac{\Pi}{\tau}, (21)
dΠdτ+ΠτΠ\displaystyle\frac{d\Pi}{d\tau}+\frac{\Pi}{\tau_{\Pi}} =\displaystyle= βΠτδΠΠΠτ+λΠππτ;\displaystyle-\frac{\beta_{\Pi}}{\tau}-\delta_{\Pi\Pi}\frac{\Pi}{\tau}+\lambda_{\Pi\pi}\frac{\pi}{\tau}; (22)

where π=π00πzz\pi=\pi^{00}-\pi^{zz}. We note that, under Björken expansion, ωμν=0\omega^{\mu\nu}=0 which implies that the term πϕμωνϕ\pi_{\phi}^{\langle\mu}\omega^{\nu\rangle\phi} in Eq. (18) vanishes and has no impact on the evolution of QGP. Temperature dependence of the shear and bulk second order transport coefficients appearing in the above equations is analyzed in Ref. Bhadury et al. (2021) and their analysis indicate that the thermal QCD medium effects have significant impact on these coefficients. Note that, Eqs. (21) and (22) give the evolution for π\pi and Π\Pi respectively governed by their corresponding relaxation times. These equations, coupled with Eq. (20) can be solved numerically by fixing τπ,τΠ\tau_{\pi},\tau_{\Pi}. In the present work, in order to study the effect of relaxation time on evolution and subsequently on signals, we choose different temperature dependent values lying between τR=2(2ln2)T(η/s)\tau_{R}=\frac{2(2-\ln 2)}{T}(\eta/s), the result corresponding to N=4N=4 supersymmetric Yang-Mills theory Baier et al. (2008); Natsuume and Okamura (2008) and τR=5.05.9T(η/s)\tau_{R}=\frac{5.0...5.9}{T}(\eta/s), which is motivated by kinetic theory York and Moore (2009). We take the following forms of relaxation time in our analysis : τπ=b1(η/s)/T\tau_{\pi}=b_{1}(\eta/s)/T and τΠ=b2(ζ/s)/T\tau_{\Pi}=b_{2}(\zeta/s)/T. The value of η/s\eta/s is taken to be 1/4π1/4\pi Kovtun et al. (2005) and ζ/s\zeta/s as temperature dependent according to the studies of strongly interacting gauge theories as given below Kanitscheider and Skenderis (2009)

ζs=2ηs(13cs2)κ(T)ηs.\frac{\zeta}{s}=2\frac{\eta}{s}\left(\frac{1}{3}-c_{s}^{2}\right)\equiv\kappa(T)\frac{\eta}{s}. (23)

As mentioned earlier, due to RTA, we obtain a single relaxation time-scale for both shear and bulk i.e.,τπ=τΠ=τRi.e.,\tau_{\pi}=\tau_{\Pi}=\tau_{R}. This condition together with Eq. (23) would give the relation b2=b1/κ(T)b_{2}=b_{1}/\kappa(T), which ensures that shear and bulk relaxation times are equal for any value of b1b_{1}. Following the analysis in Ref. Bhadury et al. (2021), we also consider a constant temperature independent relaxation time, τR=0.15\tau_{R}=0.15 fm/c in our study.

Now, we investigate the sensitivity of dissipative quantities to the temperature dependence of τR\tau_{R}. We solve Eqs. (20), (21) and (22) numerically by providing the initial conditions relevant to RHIC energies. We choose the initial proper time and temperature to be τ0=0.5\tau_{0}=0.5 fm/c and T0=0.31T_{0}=0.31 GeV respectively, following the previous studies Srivastava (1999); Bhatt et al. (2010); Chandra and Sreekanth (2015). The initial values of viscous contributions are taken as π(τ0)=Π(τ0)=0\pi(\tau_{0})=\Pi(\tau_{0})=0 GeV/fm3. The (2+1) flavor lattice QCD EoS Cheng et al. (2008); Borsanyi et al. (2014) is used to close this system of equations. In Figs. 1 and 2, the proper time evolution of shear stress tensor and magnitude of bulk viscous pressure respectively are plotted for different temperature dependent τR\tau_{R}. The evolution is plotted for τR=2(η/s)/T, 1.5(η/s)/T\tau_{R}=2(\eta/s)/T,\,1.5(\eta/s)/T and (η/s)/T(\eta/s)/T. We observe that both shear and bulk pressures have strong dependency on the form of τR\tau_{R}. The effect of viscous contributions are high for τR=2(η/s)/T\tau_{R}=2(\eta/s)/T, while it is the lowest for τR=(η/s)/T\tau_{R}=(\eta/s)/T. We also plot the viscous terms for a constant value of relaxation time, τR=0.15\tau_{R}=0.15 fm/c, chosen arbitrarily. We note that, at early times, the shear and bulk pressures are found to be high compared to the later times for every τR\tau_{R} considered here.

Refer to caption
Figure 1: Proper time evolution of shear stress tensor for different temperature dependent forms of τR\tau_{R}. Initial conditions are taken to be T0=0.31T_{0}=0.31 GeV and τ0=0.5\tau_{0}=0.5 fm/c. Evolution corresponding to τR=0.15\tau_{R}=0.15 fm/c is also plotted for comparison.
Refer to caption
Figure 2: Proper time evolution of bulk viscous pressure for different temperature dependent forms of τR\tau_{R}. Initial conditions are taken to be T0=0.31T_{0}=0.31 GeV and τ0=0.5\tau_{0}=0.5 fm/c. Evolution corresponding to τR=0.15\tau_{R}=0.15 fm/c is also plotted for comparison.

In Fig. 3, we analyze the pressure anisotropy of the medium, which is defined as

PL/PTP+ΠπP+Π+π/2,P_{L}/P_{T}\equiv\frac{P+\Pi-\pi}{P+\Pi+\pi/2}, (24)

for the various temperature forms of τR\tau_{R}. Here PL,PTP_{L},P_{T} refer to the longitudinal, transverse pressures respectively. It can be seen that, as τR\tau_{R} increases to 3(η/s)/T3(\eta/s)/T, the ratio PL/PTP_{L}/P_{T} becomes negative causing cavitation in the medium and this stops the validity of hydrodynamics. For τR=3(η/s)/T\tau_{R}=3(\eta/s)/T, we plot the curve only till pressure anisotropy approaches zero and it is observed that this occurs within a very small time of 11 fm/c. In the same figure, we plot PL/PTP_{L}/P_{T} for the constant value, τR=0.15\tau_{R}=0.15 fm/c and it is observed that cavitation scenario is not present for this particular case for the initial condition considered. However, cavitation can arise in the system even for very small values of temperature independent relaxation times greater than 0.150.15 fm/c. For instance, we can observe cavitation for τR=0.25\tau_{R}=0.25 fm/c (value taken in Ref. Bhadury et al. (2021)), around 0.0. fm/c for the initial conditions taken in this work.

Refer to caption
Figure 3: Proper time evolution of pressure anisotropy PL/PTP_{L}/P_{T} with various temperature dependent relaxation times. Initial conditions are taken to be T0=0.31T_{0}=0.31 GeV and τ0=0.5\tau_{0}=0.5 fm/c. Evolution for constant τR\tau_{R} is also shown for comparison.
Refer to caption
Figure 4: Proper time evolution of dT/dτdT/d\tau by varying the relaxation time.

We now turn our attention towards the limiting values of relaxation times possible in such an application of the formalism to heavy ion collisions by looking to the rate of temperature variation of the fireball as it evolves. We plot the rate of change of temperature as a function of proper time for different τR\tau_{R} as shown in Fig. 4. We observe that the slope of temperature evolution crosses the line T[τ]=0T^{\prime}[\tau]=0 as we increase the value of τR\tau_{R}. This indicates that, when τR\tau_{R} is large, the system undergoes reheating as it expands. Reheating of the fireball is an unphysical scenario in relativistic heavy ion collisions, since it is expected that the temperature of QGP has to decrease monotonically during its expansion Muronga (2002); Baier et al. (2006). This reheating can be observed for τR>8(η/s)/T\tau_{R}>8(\eta/s)/T and for the constant values greater than 0.50.5 fm/c. It can be seen that, for τR=8(η/s)/T\tau_{R}=8(\eta/s)/T and 0.50.5 fm/c, T[τ]T^{\prime}[\tau] crosses zero around τ=2\tau=2 fm/c.

Refer to caption
Figure 5: Comparison of evolution of shear and bulk viscous pressures obtained within EQPM with that obtained using the standard hydrodynamics taken from Ref. Muronga (2004).

Next, we compare the evolution of viscous quantities obtained within EQPM with another standard hydrodynamic framework. We take the second order dissipative hydrodynamics described in Ref. Muronga (2004) for this analysis. Within this framework, the evolution equations for viscous quantities (π\pi and Π\Pi) in Björken flow are given by

dπdτ\displaystyle\frac{d\pi}{d\tau} =\displaystyle= πτππ2(1τ+1β2Tddτ(β2T))+231β21τ,\displaystyle-\frac{\pi}{\tau_{\pi}}-\frac{\pi}{2}\left(\frac{1}{\tau}+\frac{1}{\beta_{2}}T\frac{d}{d\tau}\left(\frac{\beta_{2}}{T}\right)\right)+\frac{2}{3}\frac{1}{\beta_{2}}\frac{1}{\tau}, (25)
dΠdτ\displaystyle\frac{d\Pi}{d\tau} =\displaystyle= ΠτΠ12Πβ0(β0τ+Tddτ(β0T))1β01τ,\displaystyle-\frac{\Pi}{\tau_{\Pi}}-\frac{1}{2}\frac{\Pi}{\beta_{0}}\left(\frac{\beta_{0}}{\tau}+T\frac{d}{d\tau}\left(\frac{\beta_{0}}{T}\right)\right)\frac{1}{\beta_{0}}\frac{1}{\tau}, (26)

where β0\beta_{0} and β2\beta_{2} are related to the relaxation times as τΠ=ζβ0\tau_{\Pi}=\zeta\beta_{0} and τπ=2ηβ2\tau_{\pi}=2\eta\beta_{2}. Note that, in our analysis, we have τπ=τΠ=τR\tau_{\pi}=\tau_{\Pi}=\tau_{R}. In Fig. 5, we compare the evolution of shear and bulk viscous pressures within EQPM with that obtained using the above hydrodynamic framework. Equations (25) and (26) are numerically solved together with the evolution equation for energy density (Eq. (20)) by providing the lattice EoS Cheng et al. (2008); Borsanyi et al. (2014). We adopt the same initial conditions as before and also take τR=(η/s)/T\tau_{R}=(\eta/s)/T in our calculations. The shear stress evolution in standard hydrodynamics is observed to be large compared to that of EQPM; while, the evolution of bulk pressure is high in the EQPM framework. Moreover, we note that for τR=3(η/s)/T\tau_{R}=3(\eta/s)/T, the pressure anisotropy does not become negative within this hydrodynamic description compared to the case of EQPM.

III Thermal dilepton and photon production rates

Dissipation in the system influences the particle production rates in two ways: firstly through the hydrodynamic evolution of the system and secondly via non-equilibrium corrections to the single particle distribution functions. In the previous section, we have analyzed the impact of viscosity on the evolution of QGP. Now, we incorporate the effect of viscosities on thermal dilepton and photon production rates through viscous modified distribution functions upto first order in gradients. The major source of thermal dileptons in QGP medium is from the qq¯q\bar{q}-annihilation, qq¯γl+lq\bar{q}\rightarrow\gamma^{*}\rightarrow l^{+}l^{-}. From relativistic kinetic theory, the rate of dilepton production for this process, within the EQPM can be written as

dNd4xd4p\displaystyle\frac{dN}{d^{4}xd^{4}p} =\displaystyle= d3p1(2π)3d3p2(2π)3Meff2g2σ(Meff2)2ω1ω2\displaystyle\iint\frac{d^{3}\vec{p}_{1}}{(2\pi)^{3}}\frac{d^{3}\vec{p}_{2}}{(2\pi)^{3}}\,\frac{M_{\textrm{eff}}^{2}\,g^{2}\,\sigma(M_{\textrm{eff}}^{2})}{2\omega_{1}\omega_{2}} (27)
×fq(p1)fq(p2)δ4(p~p~1p~2).\displaystyle\times f_{q}(\vec{p}_{1})f_{q}(\vec{p}_{2})\delta^{4}(\tilde{p}-\tilde{p}_{1}-\tilde{p}_{2}).

Here, p~1,2=(ω1,2,p1,2)\tilde{p}_{1,2}=(\omega_{1,2},\vec{p}_{1,2}) is the four-momentum of the quark, anti-quark respectively, with ω1,2\omega_{1,2} being the corresponding modified single particle energy as given by Eq. (3). When quark (anti-quark) masses are neglected, we can write ω1,2|p|\omega_{1,2}\approx|\vec{p}|. Four-momentum of the dilepton is given as p~=(ω0=ω1+ω2,p=p1+p2)\tilde{p}=(\omega_{0}=\omega_{1}+\omega_{2},\vec{p}=\vec{p}_{1}+\vec{p}_{2}). The quantity Meff2=(ω1+ω2)2(p1+p2)2M_{\textrm{eff}}^{2}=(\omega_{1}+\omega_{2})^{2}-(\vec{p}_{1}+\vec{p}_{2})^{2} represents the modified effective mass of the virtual photon in the interacting QCD medium. Keeping the terms up to linear order in δωq\delta\omega_{q}, we get Naik et al. (2021a)

Meff2\displaystyle M_{\textrm{eff}}^{2} \displaystyle\approx M2(1+4δωq(E1+E2)M2),\displaystyle M^{2}\left(1+\frac{4\,\delta\omega_{q}\,(E_{1}+E_{2})}{M^{2}}\right), (28)

where M2M^{2} is the invariant mass of dilepton in the ultrarelativistic limit, zq1z_{q}\rightarrow 1. In Eq. (27), the term σ(Meff2)\sigma(M_{\textrm{eff}}^{2}) represents the cross-section for the qq¯q\bar{q}-annihilation process and gg is the degeneracy factor. In the Born approximation, we obtain Meff2g2σ(Meff2)=80π9αe2M_{\textrm{eff}}^{2}\,g^{2}\,\sigma(M_{\textrm{eff}}^{2})=\frac{80\pi}{9}\alpha_{e}^{2} Alam et al. (1996) for Nf=2N_{f}=2 and Nc=3N_{c}=3, with αe\alpha_{e} being the electromagnetic coupling constant.

We note that, in Eq. (27), fq(p)fq0+fq0f¯q0δfqf_{q}(\vec{p})\equiv f_{q}^{0}+f_{q}^{0}\bar{f}_{q}^{0}\delta f_{q} represents the quark (anti-quark) momentum distribution function away from equilibrium, with the form of equilibrium distribution function fq0f_{q}^{0} as given by Eq. (1). The viscous modification to the distribution function (Eq. (13)) can be rewritten in terms of first order gradients of hydrodynamic quantities by employing Eq. (14) and the form of δfq\delta f_{q} thus obtained is given below:

δfq\displaystyle\delta f_{q} =\displaystyle= δfπ+δfΠ\displaystyle\delta f_{\pi}+\delta f_{\Pi} (29)
=\displaystyle= β2βπ(up~)p~μp~νπμν+βΠβΠ[ξ1ξ2(up~)],\displaystyle\frac{\beta}{2\beta_{\pi}(u\cdot\tilde{p})}\tilde{p}^{\mu}\tilde{p}^{\nu}\pi_{\mu\nu}+\frac{\beta\Pi}{\beta_{\Pi}}\Big{[}\xi_{1}-\xi_{2}(u\cdot\tilde{p})\Big{]},

where

ξ1\displaystyle\xi_{1} =\displaystyle= βcs2δωqβ+δωq,\displaystyle\beta c_{s}^{2}\frac{\partial\delta\omega_{q}}{\partial\beta}+\delta\omega_{q},
ξ2\displaystyle\xi_{2} =\displaystyle= (cs213)+δωq3(up~)2[2(up~)δωq].\displaystyle\left(c_{s}^{2}-\frac{1}{3}\right)+\frac{\delta\omega_{q}}{3(u\cdot\tilde{p})^{2}}\left[2(u\cdot\tilde{p})-\delta\omega_{q}\right]. (30)

We intend to calculate the spectra of dileptons with large invariant mass i.e., M>>TM>>T. Hence we can approximate the distribution function with the Maxwell-Boltzmann one, f(p)zqeω/Tf(\vec{p})\approx z_{q}e^{-\omega/T}. Keeping the terms up to second order in momenta, we write the total dilepton production rate as Bhatt et al. (2012)

dNd4xd4p=dN(0)d4xd4p+dN(π)d4xd4p+dN(Π)d4xd4p.\frac{dN}{d^{4}xd^{4}p}=\frac{dN^{(0)}}{d^{4}xd^{4}p}+\frac{dN^{(\pi)}}{d^{4}xd^{4}p}+\frac{dN^{(\Pi)}}{d^{4}xd^{4}p}. (31)

The ideal part of the rate, within the EQPM is obtained as Chandra and Sreekanth (2015); Naik et al. (2021a)

dN(0)d4xd4p\displaystyle\frac{dN^{(0)}}{d^{4}xd^{4}p} =\displaystyle= zq2d3p1(2π)6Meff2g2σ(Meff2)2ω1ω2e(ω1+ω2)/T\displaystyle z_{q}^{2}\int\frac{d^{3}\vec{p}_{1}}{(2\pi)^{6}}\,\frac{M_{\textrm{eff}}^{2}\,g^{2}\,\sigma(M_{\textrm{eff}}^{2})}{2\omega_{1}\omega_{2}}\,e^{-(\omega_{1}+\omega_{2})/T} (32)
×δ(ω0ω1ω2)\displaystyle\times\delta(\omega_{0}-\omega_{1}-\omega_{2})
=\displaystyle= zq22Meff2g2σ(Meff2)(2π)5eω0/T.\displaystyle\frac{z_{q}^{2}}{2}\frac{M_{\textrm{eff}}^{2}\,g^{2}\,\sigma(M_{\textrm{eff}}^{2})}{(2\pi)^{5}}e^{-\omega_{0}/T}.

The contribution to the thermal dilepton rate due to shear viscosity can be written as

dN(π)d4xd4p\displaystyle\frac{dN^{(\pi)}}{d^{4}xd^{4}p} =\displaystyle= zq2d3p1(2π)3d3p2(2π)3e(ω1+ω2)/T\displaystyle z_{q}^{2}\iint\frac{d^{3}\vec{p}_{1}}{(2\pi)^{3}}\frac{d^{3}\vec{p}_{2}}{(2\pi)^{3}}\,e^{-(\omega_{1}+\omega_{2})/T} (33)
×[ββπ(up~)]Meff2g2σ(Meff2)2ω1ω2\displaystyle\times\left[\frac{\beta}{\beta_{\pi}(u\cdot\tilde{p})}\right]\frac{M_{\textrm{eff}}^{2}\,g^{2}\,\sigma(M_{\textrm{eff}}^{2})}{2\omega_{1}\omega_{2}}
×δ4(p~p~1p~2)p~1μp~1νπμν.\displaystyle\times\delta^{4}(\tilde{p}-\tilde{p}_{1}-\tilde{p}_{2})\,\,\tilde{p}_{1}^{\mu}\tilde{p}_{1}^{\nu}\pi_{\mu\nu}.

We note that, shear viscous correction to distribution function, δfπ\delta f_{\pi} is qualitatively similar to the non-equilibrium correction obtained in Refs. Bhalerao et al. (2013); Naik et al. (2021a, b). Hence, by following the analysis of the above Refs., we obtain the contribution to the dilepton rate due to shear viscosity as

dN(π)d4xd4p\displaystyle\frac{dN^{(\pi)}}{d^{4}xd^{4}p} =\displaystyle= dN(0)d4xd4p{ββπ12|p|5[(up~)|p|2(2|p|23Meff2)\displaystyle\frac{dN^{(0)}}{d^{4}xd^{4}p}\Bigg{\{}\frac{\beta}{\beta_{\pi}}\frac{1}{2|\vec{p}|^{5}}\Bigg{[}\frac{(u\cdot\tilde{p})|\vec{p}|}{2}\left(2|\vec{p}|^{2}-3M_{\textrm{eff}}^{2}\right) (34)
+34Meff4ln((up~)+|p|(up~)|p|)]p~μp~νπμν},\displaystyle+\frac{3}{4}M_{\textrm{eff}}^{4}\ln\left(\frac{(u\cdot\tilde{p})+|\vec{p}|}{(u\cdot\tilde{p})-|\vec{p}|}\right)\Bigg{]}\tilde{p}^{\mu}\tilde{p}^{\nu}\pi_{\mu\nu}\Bigg{\}},

where |p|=(up~)2Meff2|\vec{p}|=\sqrt{(u\cdot\tilde{p})^{2}-M_{\textrm{eff}}^{2}}. Following the same analysis, we calculate the contribution to the dilepton rate due to bulk viscosity as

dN(Π)d4xd4p\displaystyle\frac{dN^{(\Pi)}}{d^{4}xd^{4}p} =\displaystyle= dN(0)d4xd4p2βΠβΠ{βcs2δωqβ23δωq(cs213)(up~)2\displaystyle\frac{dN^{(0)}}{d^{4}xd^{4}p}\frac{2\beta\Pi}{\beta_{\Pi}}\Bigg{\{}\beta c_{s}^{2}\frac{\partial\delta\omega_{q}}{\partial\beta}-\frac{2}{3}\delta\omega_{q}-\left(c_{s}^{2}-\frac{1}{3}\right)\frac{(u\cdot\tilde{p})}{2} (35)
+δωq2312|p|5[(up~)|p|2(2|p|23Meff2)\displaystyle+\frac{\delta\omega_{q}^{2}}{3}\frac{1}{2|\vec{p}|^{5}}\Bigg{[}\frac{(u\cdot\tilde{p})|\vec{p}|}{2}\left(2|\vec{p}|^{2}-3M_{\textrm{eff}}^{2}\right)
+34Meff4ln((up~)+|p|(up~)|p|)]}.\displaystyle+\frac{3}{4}M_{\textrm{eff}}^{4}\ln\left(\frac{(u\cdot\tilde{p})+|\vec{p}|}{(u\cdot\tilde{p})-|\vec{p}|}\right)\Bigg{]}\Bigg{\}}.

We now determine the modification to the thermal photon emission rate due to shear and bulk viscosities. We consider two major sources of thermal photons: Compton scattering, q(q¯)gq(q¯)γq(\bar{q})g\rightarrow q(\bar{q})\gamma and qq¯q\bar{q}-annihilation, qq¯gγq\bar{q}\rightarrow g\gamma. We emphasize that, in the present work we only examine the hard contributions of thermal photon emission. Following Bhatt et al. (2010); Dusling (2010); Wong (1995), the total photon rate in the presence of dissipation within EQPM can be written as

ω0dNγd3pd4x\displaystyle\omega_{0}\frac{dN_{\gamma}}{d^{3}pd^{4}x} =\displaystyle= 59αeαs2π2T2fq(p)[ln(12(up~)g2T)+Cann+CComp2]\displaystyle\frac{5}{9}\frac{\alpha_{e}\alpha_{s}}{2\pi^{2}}T^{2}f_{q}(\vec{p})\left[\ln\left(\frac{12\,(u\cdot\tilde{p})}{g^{2}T}\right)\!+\!\frac{C_{\textrm{ann}}\!+\!C_{\textrm{Comp}}}{2}\right] (36)
=\displaystyle= ω0dNγ(0)d3pd4x+ω0dNγ(π)d3pd4x+ω0dNγ(Π)d3pd4x,\displaystyle\omega_{0}\frac{dN_{\gamma}^{(0)}}{d^{3}pd^{4}x}+\omega_{0}\frac{dN_{\gamma}^{(\pi)}}{d^{3}pd^{4}x}+\omega_{0}\frac{dN_{\gamma}^{(\Pi)}}{d^{3}pd^{4}x},

where the constants take the values Cann=1.91613C_{\textrm{ann}}=-1.91613, CComp=0.41613C_{\textrm{Comp}}=-0.41613. Here αs\alpha_{s} denotes the strong coupling constant and g=4παsg=\sqrt{4\pi\alpha_{s}}. The ideal contribution to the photon rate takes the following form:

ω0dNγ(0)d3pd4x=59αeαs2π2T2zq2e(up~)/Tln(3.7388(up~)g2T).\omega_{0}\frac{dN_{\gamma}^{(0)}}{d^{3}pd^{4}x}=\frac{5}{9}\frac{\alpha_{e}\alpha_{s}}{2\pi^{2}}T^{2}z_{q}^{2}e^{-(u\cdot\tilde{p})/T}\ln\left(\frac{3.7388\,(u\cdot\tilde{p})}{g^{2}T}\right). (37)

Viscous contributions to photon rate are obtained as

ω0dNγ(π)d3pd4x\displaystyle\omega_{0}\frac{dN_{\gamma}^{(\pi)}}{d^{3}pd^{4}x} =\displaystyle= ω0dNγ(0)d3pd4x{β2βπ(up~)},\displaystyle\omega_{0}\frac{dN_{\gamma}^{(0)}}{d^{3}pd^{4}x}\left\{\frac{\beta}{2\beta_{\pi}(u\cdot\tilde{p})}\right\}, (38)
ω0dNγ(Π)d3pd4x\displaystyle\omega_{0}\frac{dN_{\gamma}^{(\Pi)}}{d^{3}pd^{4}x} =\displaystyle= ω0dNγ(0)d3pd4x{βΠβΠ[ξ1ξ2(up~)]},\displaystyle\omega_{0}\frac{dN_{\gamma}^{(0)}}{d^{3}pd^{4}x}\left\{\frac{\beta\Pi}{\beta_{\Pi}}\Big{[}\xi_{1}-\xi_{2}(u\cdot\tilde{p})\Big{]}\right\}, (39)

where ξ1\xi_{1} and ξ2\xi_{2} are defined in Eq. (III).

Next, in order to do a comparative study with the standard hydrodynamic results, we calculate the thermal dilepton and photon emission rates by employing the viscous modified distribution function of the form: f=f0(1+δϕ)f=f_{0}(1+\delta\phi), where f0=eE/Tf_{0}=e^{-E/T}. The non-equilibrium correction is obtained from the 1414-moment Grad’s method and is shown below Dusling and Teaney (2008) :

δϕ=pμpν2(ϵ+P)T2(πμν+25ΠΔμν).\displaystyle\delta\phi=\frac{p^{\mu}p^{\nu}}{2(\epsilon+P)T^{2}}\left(\pi_{\mu\nu}+\frac{2}{5}\Pi\Delta_{\mu\nu}\right). (40)

The ideal part of the dilepton rate, in the absence of viscosity is well known and is given by Vogt (2007)

dN0d4xd4p=12M2g2σ(M2)(2π)5e(up)/T.\frac{dN_{0}}{d^{4}xd^{4}p}=\frac{1}{2}\frac{M^{2}\,g^{2}\,\sigma(M^{2})}{(2\pi)^{5}}e^{-(u\cdot p)/T}. (41)

The contribution to the dilepton rate due to shear and bulk viscous terms in δϕ\delta\phi are calculated as Dusling and Lin (2008); Bhatt et al. (2012)

dNπd4xd4p\displaystyle\frac{dN_{\pi}}{d^{4}xd^{4}p} =\displaystyle= 23(pμpν2sT3πμν)dN0d4xd4p,\displaystyle\frac{2}{3}\left(\frac{p^{\mu}p^{\nu}}{2sT^{3}}\pi_{\mu\nu}\right)\frac{dN_{0}}{d^{4}xd^{4}p}, (42)
dNΠd4xd4p\displaystyle\frac{dN_{\Pi}}{d^{4}xd^{4}p} =\displaystyle= 25sT3(M212gαβ13pαpβ)ΔαβΠdN0d4xd4p,\displaystyle\frac{2}{5sT^{3}}\left(\frac{M^{2}}{12}g^{\alpha\beta}-\frac{1}{3}p^{\alpha}p^{\beta}\right)\Delta_{\alpha\beta}\Pi\frac{dN_{0}}{d^{4}xd^{4}p},

respectively. Similarly, the ideal and viscous contributions to thermal photon rate are obtained as given below Dusling (2010); Bhatt et al. (2010) :

EdN0d3pd4x\displaystyle E\frac{dN_{0}}{d^{3}pd^{4}x} =\displaystyle= 59αeαs2π2T2e(up)/Tln(3.7388(up)g2T),\displaystyle\frac{5}{9}\frac{\alpha_{e}\alpha_{s}}{2\pi^{2}}T^{2}e^{-(u\cdot p)/T}\ln\left(\frac{3.7388\,(u\cdot p)}{g^{2}T}\right), (44)
EdNπd3pd4x\displaystyle E\frac{dN_{\pi}}{d^{3}pd^{4}x} =\displaystyle= EdN0d3pd4x{pμpνπμν2(ϵ+P)T2},\displaystyle E\frac{dN_{0}}{d^{3}pd^{4}x}\left\{\frac{p^{\mu}p^{\nu}\pi_{\mu\nu}}{2(\epsilon+P)T^{2}}\right\}, (45)
EdNΠd3pd4x\displaystyle E\frac{dN_{\Pi}}{d^{3}pd^{4}x} =\displaystyle= EdN0d3pd4x{25ΠpμpνΔμν2(ϵ+P)T2}.\displaystyle E\frac{dN_{0}}{d^{3}pd^{4}x}\left\{\frac{2}{5}\Pi\frac{p^{\mu}p^{\nu}\Delta_{\mu\nu}}{2(\epsilon+P)T^{2}}\right\}. (46)

In the next section, we proceed to calculate the thermal dilepton and photon spectra from heavy-ion collisions by convoluting the rate expressions over the space-time evolution of the collisions along with the temperature profile and viscous evolution of the QGP obtained in section II.1.

IV Particle spectra from an expanding QGP

Thermal dileptons and photons are produced from a thermalized QGP throughout its evolution. We calculate the emission spectra of these particles over the entire evolution by considering Björken’s 11-D model Bjorken (1983). Within the model, four-dimensional volume element in the Minkowski space gets modified as d4x=Aτdτdηsd^{4}x=A_{\bot}\tau d\tau d\eta_{s}. Here, A=πRA2A_{\bot}=\pi R_{A}^{2} is the transverse area of the collision, with RA=1.2A1/3R_{A}=1.2\,A^{1/3} fm being the radius of the colliding nuclei (for Au, A=197A=197). Now, we write the total thermal dilepton and photon yields in terms of transverse momentum pTp_{T}, invariant mass MM and rapidity yy of the particle produced

dNdM2d2pTdy\displaystyle\frac{dN}{dM^{2}d^{2}p_{T}dy} =\displaystyle= Aτ0τf𝑑ττ𝑑ηsχ(T,ηs)(12dNd4xd4p)\displaystyle A_{\bot}\,\int_{\tau_{0}}^{\tau_{f}}d\tau\,\tau\int_{-\infty}^{\infty}d\eta_{s}\,\chi(T,\eta_{s})\left(\frac{1}{2}\frac{dN}{d^{4}xd^{4}p}\right)
dNd2pTdy\displaystyle\frac{dN}{d^{2}p_{T}dy} =\displaystyle= Aτ0τf𝑑ττ𝑑ηs(ωdNd3pd4x),\displaystyle A_{\bot}\,\int_{\tau_{0}}^{\tau_{f}}d\tau\,\tau\int_{-\infty}^{\infty}d\eta_{s}\,\left(\omega\frac{dN}{d^{3}pd^{4}x}\right), (47)

where the factor χ(T,ηs)=[1+2mTcosh(yηs)δωq]\chi(T,\eta_{s})=\left[1+\frac{2}{m_{T}}\cosh(y-\eta_{s})\delta\omega_{q}\right] and τf\tau_{f} is the time taken by the fireball to cool down to the critical temperature TcT_{c}. Four-momentum of the dilepton can be parametrized as p~μ=(MTcoshy,pTcosϕp,pTsinϕp,MTsinhy)\tilde{p}^{\mu}=(M_{T}\cosh y,p_{T}\cos\phi_{p},p_{T}\sin\phi_{p},M_{T}\sinh y), where MT=pT2+Meff2M_{T}=\sqrt{p_{T}^{2}+M_{\textrm{eff}}^{2}} is the medium modified transverse mass of dilepton. Now, for an expanding QGP within Björken flow, we evaluate the factors appearing in particle rate as

(up~)\displaystyle(u\cdot\tilde{p}) =MTcosh(yηs),\displaystyle=M_{T}\cosh(y-\eta_{s}), (48)
p~μp~νπμν\displaystyle\tilde{p}^{\mu}\tilde{p}^{\nu}\pi_{\mu\nu} =π[pT22MT2sinh2(yηs)],\displaystyle=\pi\left[\frac{p_{T}^{2}}{2}-M_{T}^{2}\sinh^{2}(y-\eta_{s})\right], (49)
p~μp~νΔμν\displaystyle\tilde{p}^{\mu}\tilde{p}^{\nu}\Delta_{\mu\nu} =pT2mT2sinh2(yηs).\displaystyle=-p_{T}^{2}-m_{T}^{2}\sinh^{2}(y-\eta_{s}). (50)

By noting the above expressions, we write the ideal contribution to thermal dilepton yields as

dN(0)dM2d2pTdy\displaystyle\frac{dN^{(0)}}{dM^{2}d^{2}p_{T}dy} =\displaystyle= 𝒬τ0τf𝑑ττzq2𝑑ηsχ(T,ηs)\displaystyle\mathcal{Q}\int_{\tau_{0}}^{\tau_{f}}d\tau\,\tau z_{q}^{2}\,\int_{-\infty}^{\infty}d\eta_{s}\,\chi(T,\eta_{s}) (51)
×Exp(MTTcosh(yηs)).\displaystyle\times\textrm{Exp}\left(-\frac{M_{T}}{T}\cosh(y-\eta_{s})\right).

The viscous contributions to the dilepton yields are obtained as

dN(π)dM2d2pTdy\displaystyle\frac{dN^{(\pi)}}{dM^{2}d^{2}p_{T}dy} =\displaystyle= 𝒬τ0τf𝑑ττzq2βπ4βπ𝑑ηsχ(T,ηs)Exp(MTTcosh(yηs))[pT2/2MT2sinh2(yηs)][MT2cosh2(yηs)Meff2]5/2\displaystyle\mathcal{Q}\int_{\tau_{0}}^{\tau_{f}}d\tau\,\tau z_{q}^{2}\,\frac{\beta\pi}{4\beta_{\pi}}\int_{-\infty}^{\infty}d\eta_{s}\,\chi(T,\eta_{s})\textrm{Exp}\left(-\frac{M_{T}}{T}\cosh(y-\eta_{s})\right)\frac{\left[p_{T}^{2}/2-M_{T}^{2}\sinh^{2}(y-\eta_{s})\right]}{[M_{T}^{2}\cosh^{2}(y-\eta_{s})-M_{\textrm{eff}}^{2}]^{5/2}} (52)
×{MTcosh(yηs)MT2cosh2(yηs)Meff2(2MT2cosh2(yηs)5Meff2)\displaystyle\times\Bigg{\{}M_{T}\cosh(y-\eta_{s})\sqrt{M_{T}^{2}\cosh^{2}(y-\eta_{s})-M_{\textrm{eff}}^{2}}\Big{(}2M_{T}^{2}\cosh^{2}(y-\eta_{s})-5M_{\textrm{eff}}^{2}\Big{)}
+32Meff4ln(MTcosh(yηs)+MT2cosh2(yηs)Meff2MTcosh(yηs)MT2cosh2(yηs)Meff2)},\displaystyle+\frac{3}{2}M_{\textrm{eff}}^{4}\ln\left(\frac{M_{T}\cosh(y-\eta_{s})+\sqrt{M_{T}^{2}\cosh^{2}(y-\eta_{s})-M_{\textrm{eff}}^{2}}}{M_{T}\cosh(y-\eta_{s})-\sqrt{M_{T}^{2}\cosh^{2}(y-\eta_{s})-M_{\textrm{eff}}^{2}}}\right)\Bigg{\}},
dN(Π)dM2d2pTdy\displaystyle\frac{dN^{(\Pi)}}{dM^{2}d^{2}p_{T}dy} =\displaystyle= 𝒬τ0τf𝑑ττzq22βΠβΠ𝑑ηsχ(T,ηs)Exp(MTTcosh(yηs))\displaystyle\mathcal{Q}\int_{\tau_{0}}^{\tau_{f}}d\tau\,\tau z_{q}^{2}\frac{2\beta\Pi}{\beta_{\Pi}}\int_{-\infty}^{\infty}d\eta_{s}\,\chi(T,\eta_{s})\textrm{Exp}\left(-\frac{M_{T}}{T}\cosh(y-\eta_{s})\right) (53)
×{βcs2δωqβ23δωq12(cs213)MTcosh(yηs)+δωq2314[MT2cosh2(yηs)Meff2]5/2\displaystyle\times\Bigg{\{}\beta c_{s}^{2}\frac{\partial\delta\omega_{q}}{\partial\beta}-\frac{2}{3}\delta\omega_{q}-\frac{1}{2}\left(c_{s}^{2}-\frac{1}{3}\right)M_{T}\cosh(y-\eta_{s})+\frac{\delta\omega_{q}^{2}}{3}\frac{1}{4[M_{T}^{2}\cosh^{2}(y-\eta_{s})-M_{\textrm{eff}}^{2}]^{5/2}}
×[MTcosh(yηs)MT2cosh2(yηs)Meff2(2MT2cosh2(yηs)5Meff2)\displaystyle\times\Bigg{[}M_{T}\cosh(y-\eta_{s})\sqrt{M_{T}^{2}\cosh^{2}(y-\eta_{s})-M_{\textrm{eff}}^{2}}\Big{(}2M_{T}^{2}\cosh^{2}(y-\eta_{s})-5M_{\textrm{eff}}^{2}\Big{)}
+32Meff4ln(MTcosh(yηs)+MT2cosh2(yηs)Meff2MTcosh(yηs)MT2cosh2(yηs)Meff2)]},\displaystyle+\frac{3}{2}M_{\textrm{eff}}^{4}\ln\left(\frac{M_{T}\cosh(y-\eta_{s})+\sqrt{M_{T}^{2}\cosh^{2}(y-\eta_{s})-M_{\textrm{eff}}^{2}}}{M_{T}\cosh(y-\eta_{s})-\sqrt{M_{T}^{2}\cosh^{2}(y-\eta_{s})-M_{\textrm{eff}}^{2}}}\right)\Bigg{]}\Bigg{\}},

where 𝒬=A4(2π)580π9αe2\mathcal{Q}=\frac{A_{\bot}}{4(2\pi)^{5}}\frac{80\pi}{9}\alpha_{e}^{2}. From Eq. (31), we note the total dilepton yield to be

dNdM2d2pTdy=dN(0)dM2d2pTdy+dN(π)dM2d2pTdy+dN(Π)dM2d2pTdy.\frac{dN}{dM^{2}d^{2}p_{T}dy}=\frac{dN^{(0)}}{dM^{2}d^{2}p_{T}dy}+\frac{dN^{(\pi)}}{dM^{2}d^{2}p_{T}dy}+\frac{dN^{(\Pi)}}{dM^{2}d^{2}p_{T}dy}.

The thermal photon yield is obtained as

dNd2pTdy=dN(0)d2pTdy+dN(π)d2pTdy+dN(Π)d2pTdy.\frac{dN}{d^{2}p_{T}dy}=\frac{dN^{(0)}}{d^{2}p_{T}dy}+\frac{dN^{(\pi)}}{d^{2}p_{T}dy}+\frac{dN^{(\Pi)}}{d^{2}p_{T}dy}. (54)

Noting the photon energy to be (up~)=pTcosh(yηs)(u\cdot\tilde{p})=p_{T}\cosh(y-\eta_{s}), the ideal part of photon yield is obtained as

dN(0)d2pTdy\displaystyle\frac{dN^{(0)}}{d^{2}p_{T}dy} =\displaystyle= 𝒞τ0τf𝑑ττzq2T2𝑑ηsExp(pTTcosh(yηs))\displaystyle\mathcal{C}\int_{\tau_{0}}^{\tau_{f}}d\tau\,\tau z_{q}^{2}T^{2}\int_{-\infty}^{\infty}d\eta_{s}\textrm{Exp}\left(-\frac{p_{T}}{T}\cosh(y-\eta_{s})\right) (55)
×ln(3.7388pTcosh(yηs)g2T).\displaystyle\times\ln\left(\frac{3.7388\,p_{T}\cosh(y-\eta_{s})}{g^{2}T}\right).

Viscous contributions to photon yield are obtained as

dN(π)d2pTdy\displaystyle\frac{dN^{(\pi)}}{d^{2}p_{T}dy} =\displaystyle= 𝒞τ0τf𝑑ττzq2T2𝑑ηsExp(pTTcosh(yηs))\displaystyle\mathcal{C}\int_{\tau_{0}}^{\tau_{f}}d\tau\,\tau z_{q}^{2}T^{2}\int_{-\infty}^{\infty}d\eta_{s}\textrm{Exp}\left(-\frac{p_{T}}{T}\cosh(y-\eta_{s})\right) (56)
×β2βπ1pTcosh(yηs)ln(3.7388pTcosh(yηs)g2T)\displaystyle\times\frac{\beta}{2\beta_{\pi}}\frac{1}{p_{T}\cosh(y-\eta_{s})}\ln\left(\frac{3.7388\,p_{T}\cosh(y-\eta_{s})}{g^{2}T}\right)
×π[pT22MT2sinh2(yηs)]\displaystyle\times\pi\left[\frac{p_{T}^{2}}{2}-M_{T}^{2}\sinh^{2}(y-\eta_{s})\right]
dN(Π)d2pTdy\displaystyle\frac{dN^{(\Pi)}}{d^{2}p_{T}dy} =\displaystyle= 𝒞τ0τf𝑑ττzq2T2𝑑ηsExp(pTTcosh(yηs))\displaystyle\mathcal{C}\,\int_{\tau_{0}}^{\tau_{f}}d\tau\,\tau z_{q}^{2}T^{2}\int_{-\infty}^{\infty}d\eta_{s}\textrm{Exp}\left(-\frac{p_{T}}{T}\cosh(y-\eta_{s})\right) (57)
×βΠβΠ[ξ1ξ2pTcosh(yηs)]\displaystyle\times\frac{\beta\Pi}{\beta_{\Pi}}\left[\xi_{1}-\xi_{2}p_{T}\cosh(y-\eta_{s})\right]
×ln(3.7388pTcosh(yηs)g2T),\displaystyle\times\ln\left(\frac{3.7388\,p_{T}\cosh(y-\eta_{s})}{g^{2}T}\right),

where 𝒞=5A9αeαs2π2\mathcal{C}=\frac{5A_{\bot}}{9}\frac{\alpha_{e}\alpha_{s}}{2\pi^{2}}.

Now, we write the expressions for thermal dilepton and photon yields in the standard hydrodynamics with the non-equilibrium corrections given by Eq. (40). The ideal dilepton yield is obtained by taking the limit zq1z_{q}\rightarrow 1 and MeffMM_{eff}\rightarrow M in Eq. (51). The viscous corrections to the dilepton yields are given by

dNπdM2d2pTdy\displaystyle\frac{dN_{\pi}}{dM^{2}d^{2}p_{T}dy} =\displaystyle= 𝒬τ0τf𝑑ττ𝑑ηsExp(mTTcosh(yηs))\displaystyle\mathcal{Q}\int_{\tau_{0}}^{\tau_{f}}d\tau\,\tau\int_{-\infty}^{\infty}d\eta_{s}\textrm{Exp}\left(-\frac{m_{T}}{T}\cosh(y-\eta_{s})\right)
×π3sT3[pT22mT2sinh2(yηs)],\displaystyle\times\frac{\pi}{3sT^{3}}\left[\frac{p_{T}^{2}}{2}-m_{T}^{2}\sinh^{2}(y-\eta_{s})\right],
dNΠdM2d2pTdy\displaystyle\frac{dN_{\Pi}}{dM^{2}d^{2}p_{T}dy} =\displaystyle= 𝒬τ0τf𝑑ττ𝑑ηsExp(mTTcosh(yηs))\displaystyle\mathcal{Q}\int_{\tau_{0}}^{\tau_{f}}d\tau\,\tau\int_{-\infty}^{\infty}d\eta_{s}\textrm{Exp}\left(-\frac{m_{T}}{T}\cosh(y-\eta_{s})\right)
×2Π5sT3[M24+13(pT2+mT2sinh2(yηs))].\displaystyle\times\frac{2\Pi}{5sT^{3}}\left[\frac{M^{2}}{4}+\frac{1}{3}\Bigg{(}p_{T}^{2}+m_{T}^{2}\sinh^{2}(y-\eta_{s})\Bigg{)}\right].

The viscous corrections to the thermal photon yields are obtained as

dNπd2pTdy\displaystyle\frac{dN_{\pi}}{d^{2}p_{T}dy} =\displaystyle= 𝒞τ0τf𝑑ττT2𝑑ηsExp(pTTcosh(yηs))\displaystyle\mathcal{C}\,\int_{\tau_{0}}^{\tau_{f}}d\tau\,\tau T^{2}\int_{-\infty}^{\infty}d\eta_{s}\textrm{Exp}\left(-\frac{p_{T}}{T}\cosh(y-\eta_{s})\right)
×ln(3.7388pTcosh(yηs)g2T)\displaystyle\times\ln\left(\frac{3.7388\,p_{T}\cosh(y-\eta_{s})}{g^{2}T}\right)
×π2sT3[pT22pT2sinh2(yηs)],\displaystyle\times\frac{\pi}{2sT^{3}}\left[\frac{p_{T}^{2}}{2}-p_{T}^{2}\sinh^{2}(y-\eta_{s})\right],
dN(Π)d2pTdy\displaystyle\frac{dN^{(\Pi)}}{d^{2}p_{T}dy} =\displaystyle= 𝒞τ0τf𝑑ττT2𝑑ηsExp(pTTcosh(yηs))\displaystyle\mathcal{C}\,\int_{\tau_{0}}^{\tau_{f}}d\tau\,\tau T^{2}\int_{-\infty}^{\infty}d\eta_{s}\textrm{Exp}\left(-\frac{p_{T}}{T}\cosh(y-\eta_{s})\right) (59)
×ln(3.7388pTcosh(yηs)g2T)\displaystyle\times\ln\left(\frac{3.7388\,p_{T}\cosh(y-\eta_{s})}{g^{2}T}\right)
×Π5(ϵ+P)T2[pT2mT2sinh2(yηs)].\displaystyle\times\frac{\Pi}{5(\epsilon+P)T^{2}}\Big{[}-p_{T}^{2}-m_{T}^{2}\sinh^{2}(y-\eta_{s})\Big{]}.

V Results and Discussion

In this section, we present the particle spectra in the presence of viscous corrections for longitudinal expansion of QGP medium. The temperature and viscous profiles (T(τ),π(τ)T(\tau),\pi(\tau) and Π(τ)\Pi(\tau)) have been determined in section II.1 by numerically solving the hydrodynamic equations, Eqs. (20)-(22). In order to obtain the dilepton and photon spectra from QGP phase, we evolve the system till TcT_{c} and in this work, we fix Tc=0.17T_{c}=0.17 GeV. The value of τf\tau_{f} is determined by solving the equation : T(τ)==TcT(\tau)==T_{c}. Also, the spectra is calculated for the midrapidity region of the particles, i.e.,y=0i.e.,\,y=0.

Refer to caption
Figure 6: Thermal dilepton yields in the presence of viscous corrections corresponding to τR=0.15\tau_{R}=0.15 fm/c and for M=1M=1 GeV.
Refer to caption
Figure 7: Dilepton spectra in the presence of viscous corrections by varying τR\tau_{R} for M=1M=1 GeV
Refer to caption
Figure 8: Comparison of dilepton spectra for different MM values with τR=0.15\tau_{R}=0.15 fm/c. The solid lines represent the total yields and dashed lines correspond to δf=0\delta f=0 case.

We now study the effect of viscosity on particle spectra by comparing the total spectra with the ideal case (δf=0\delta f=0). The ideal dilepton and photon spectra is calculated by integrating the ideal contribution to the yields, Eqs. (51) and (55) along with the ideal Björken evolution, Ti(τ)=T0(τ0/τ)1/3T_{i}(\tau)=T_{0}(\tau_{0}/\tau)^{1/3}. Firstly, we analyze the impact of both shear and bulk viscous corrections to the distribution function on thermal dilepton yield. In Fig. 6, we show thermal dilepton spectra in the presence of dissipation for a constant value of relaxation time, τR=0.15\tau_{R}=0.15 fm/c. The yields are plotted for invariant mass M=1M=1 GeV. The dashed curve denotes the ideal case. We first consider the case, δf=δfπ\delta f=\delta f_{\pi}, which is computed by taking δfΠ=0\delta f_{\Pi}=0 and Π=0\Pi=0 in the analysis. We observe that shear viscosity enhances the dilepton spectra, especially at high pTp_{T}. However, this increment is observed to be marginal when compared to the yield in the presence of bulk viscosity. The yield for δfΠ\delta f_{\Pi} correction is plotted by switching off the shear contributions in Eqs. (29) and (20)-(22). Fig. 6 shows that the effect of bulk viscosity enhances the dilepton yield throughout the pTp_{T} regime. Enhancement is observed to be maximum at small pTp_{T} and minimum at large pTp_{T}. Moreover, maximum increment in yield is observed when both shear and bulk corrections (δfπ+δfΠ\delta f_{\pi}+\delta f_{\Pi}) are taken into account.

Next, we compare the thermal dilepton yields corresponding to M=1M=1 GeV for different temperature dependent τR\tau_{R} in Fig. 7. We take both δfπ\delta f_{\pi} and δfΠ\delta f_{\Pi} terms in distribution function for this comparison. It can be seen that the dilepton yield increases with the increase in magnitude of τR\tau_{R}. The maximum enhancement in the yield is observed for τR=2(η/s)/T\tau_{R}=2(\eta/s)/T and minimum for τR=(η/s)/T\tau_{R}=(\eta/s)/T. In Fig. 8, we study the impact of viscosities on dilepton yield by varying the value of MM. We plot the total yields for M=1,1.5M=1,1.5 GeV. The dashed lines indicate the corresponding yields for δf=0\delta f=0. We observe that there is an overall decrement in the yield as the value of MM increases. Also, note that enhancement to the yield due to dissipation decreases with large MM. Over the entire pTp_{T} range, enhancement due to viscosity is more for M=1M=1 GeV.

Similarly, we analyse the effect of dissipation on thermal photon yield in Figs. 9 and 10. Impact of viscous terms on photon yield is studied with τR=0.15\tau_{R}=0.15 fm/c in Fig. 9. As observed for dilepton (in Fig. 6), maximum enhancement to the yield due to bulk viscosity (δfΠ\delta f_{\Pi}) is observed at small pTp_{T}, where as increment due to shear correction (δfπ\delta f_{\pi}) is significant at high pTp_{T}. The combined effect of both (δfπ+δfΠ\delta f_{\pi}+\delta f_{\Pi}) leads to the maximum enhancement over the entire pTp_{T}. Moreover, we note that the yield in the presence of bulk viscous correction crosses the δfπ\delta f_{\pi} curve around pT1.7p_{T}\sim 1.7 GeV, where as in Fig. 6 (for dileptons), this cross-over was observed around pT2p_{T}\sim 2 GeV. Fig. 10 exhibits a comparison between thermal photon yields for different τR\tau_{R}. It is observed that enhancement in the yield is seen to be marginal with τR=(η/s)/T\tau_{R}=(\eta/s)/T and yield is greatest with τR=2(η/s)/T\tau_{R}=2(\eta/s)/T.

Finally, in Figs. 11 and 12, we compare the thermal dilepton and photon yields obtained within this new second order hydrodynamics with that of a standard one given by Eqs. (20), (25) and (26). The yields are plotted for τR=(η/s)/T\tau_{R}=(\eta/s)/T and with invariant mass M=1M=1 GeV (for dileptons). It is clear that the spectra of dileptons and photons depend strongly on the interaction effects in the medium. The presence of these interaction terms in the yield expressions within EQPM suppresses the particle production throughout the pTp_{T} regime. This is inline with the results of Ref. Chandra and Sreekanth (2017). In Fig. 11, we also plot the dilepton yield within EQPM by neglecting the interaction effects on invariant mass of dilepton i.e., in the limit MeffMM_{eff}\rightarrow M. It can be seen that the dilepton production yield increases in this limit as shown by the results of Ref. Naik et al. (2022).

Refer to caption
Figure 9: Thermal photon yield in the presence of dissipative corrections for constant relaxation time, τR=0.15\tau_{R}=0.15 fm/c.
Refer to caption
Figure 10: Comparison of thermal photon yields in the presence of total viscous correction (δfπ+δfΠ\delta f_{\pi}+\delta f_{\Pi}) by varyiing the relaxation time.
Refer to caption
Figure 11: Comparison of thermal dilepton yields obtained within EQPM with the yields calculated using the standard hydrodynamic framework in Ref. Bhalerao et al. (2013).
Refer to caption
Figure 12: Comparison of thermal photon yields obtained within EQPM with the yields calculated using the standard hydrodynamic framework in Ref. Bhalerao et al. (2013).

VI Summary

We have employed the recently developed second order dissipative hydrodynamics estimated within the effective fugacity quasiparticle model to study the thermal dilepton and photon production from QGP. In this study, we have used the non-equilibrium quark-antiquark distribution functions, up to first order in momenta, determined using the iterative CE type expansion of effective Boltzmann equation in the relaxation time approximation (RTA). The second order viscous hydrodynamic equations were solved for different temperature dependent relaxation times (τR\tau_{R}) within the 11-D boost invariant expansion of QGP and the evolution has been compared with that obtained for a temperature independent constant value τR=0.15\tau_{R}=0.15 fm/c. It must be emphasized that, the choice of this constant value is arbitrary. We have incorporated the effect of shear and bulk viscosity coefficients through their respective relaxation times τπ\tau_{\pi} and τΠ\tau_{\Pi}, ensuring τπ=τΠ\tau_{\pi}=\tau_{\Pi} as demanded by RTA. The impact of shear and bulk viscous pressures was found to increase with increment in τR\tau_{R}. By looking into the dynamical pressure anisotropy with various relaxation times, we found that cavitation scenarios can be present in the medium for small values of τR\tau_{R}. Moreover, we obtained the limiting values of relaxation times by looking into fireball reheating scenarios. Our analysis indicate that relaxation time should not increase the constant value 0.50.5 fm/c and temperature dependent value 8(1/4π)/T8(1/4\pi)/T.

Further, using this causal hydrodynamic model, we explored the thermal dilepton and photon yields from prominent production sources in the presence of viscous modified distribution functions within longitudinal expansion of the QGP. This was done after calculating the thermal particle spectra by including the viscous modified single particle distribution functions. The effect of both shear and bulk corrections under this formalism is to enhance the dilepton and photon spectra. We also analyzed the impact of dissipation on yields by varying the value of invariant mass. Thermal dilepton and photon yields were studied for different temperature dependent τR\tau_{R} and we found that maximum enhancement is observed for the largest value of τR\tau_{R} considered. Our results also indicate that thermal dilepton and photon spectra in the presence of first order dissipative terms in the distribution function does not exhibit large corrections from the ideal case and is well behaved compared to the results using Grad’s method Bhatt et al. (2012); Bhalerao et al. (2014). We also did a comparative study of the spectra calculated within EQPM hydrodynamic framework with that obtained in the absence of mean field terms, by employing a standard relativistic hydrodynamics. Our study indicates that the presence of interaction terms in the yield expressions suppress the thermal dilepton and photon spectra throughout the pTp_{T} range. In future, we would like to study thermal dilepton and photon production by employing this causal second order hydrodynamic framework in the presence of Chapman-Enskog type viscous corrections, up to second order in gradients Bhadury et al. (2021). Further, it will be interesting to do a quantitative analysis with (2+12+1)-D hydrodynamics by varying the initial conditions. We leave these aspects for future study.

Acknowledgements

L. J. N. acknowledges the Department of Science and Technology, Government of India for the INSPIRE Fellowship. We thank the anonymous referees of this article for comments which led to the improvement of quality of the manuscript.

Appendix A Second Order Transport Coefficients

The second order transport coefficients appearing in the viscous evolution Eqs. (18) and (19) are obtained as Bhadury et al. (2021)

δππ\displaystyle\delta_{\pi\pi} =53+ββπk=q,g[73J~k63(3)+δωk(73L~k63(3)76ξk+12Γk)],\displaystyle=\dfrac{5}{3}+\dfrac{\beta}{\beta_{\pi}}\sum_{k=q,g}\bigg{[}\dfrac{7}{3}\tilde{J}^{(3)}_{k~{}63}+{\delta{\omega}_{k}}\Big{(}\dfrac{7}{3}\tilde{L}^{(3)}_{k~{}63}-\dfrac{7}{6}\xi_{k}+\dfrac{1}{2}\Gamma_{k}\Big{)}\bigg{]}, (60)
τππ\displaystyle\tau_{\pi\pi} =2+ββπk=q,g[4(J~k63(3)+δωkL~k63(3))2δωkξk],\displaystyle=2+\frac{\beta}{\beta_{\pi}}\sum_{k=q,g}\bigg{[}4\Big{(}\tilde{J}^{(3)}_{k~{}63}+\delta\omega_{k}\tilde{L}^{(3)}_{k~{}63}\Big{)}-{2\delta\omega_{k}}\xi_{k}\bigg{]}, (61)
λπΠ\displaystyle\lambda_{\pi\Pi} =βcs2βΠk=q,g[J~k42(1)+J~k31(0)+δωk(L~k42(1)J~k21(0)+L~k31(0)δωkL~k21(0))+βδωkβ(2ξk+Γk\displaystyle=\dfrac{\beta c_{s}^{2}}{\beta_{\Pi}}\sum_{k=q,g}\bigg{[}\tilde{J}^{(1)}_{k~{}42}+\tilde{J}^{(0)}_{k~{}31}+\delta\omega_{k}\Big{(}\tilde{L}^{(1)}_{k~{}42}-\tilde{J}^{(0)}_{k~{}21}+\tilde{L}^{(0)}_{k~{}31}-\delta\omega_{k}\tilde{L}^{(0)}_{k~{}21}\Big{)}+\beta\dfrac{\partial\delta\omega_{k}}{\partial\beta}\Big{(}2\xi_{k}+\Gamma_{k}
2δωkL~k42(1))]+ββΠk=q,g[143J~k63(3)+103J~k42(1)+δωk(143L~k63(3)+103L~k42(1)73ξk+Γk)],\displaystyle~{}~{}-2\delta\omega_{k}\tilde{L}^{(1)}_{k~{}42}\Big{)}\!\bigg{]}+\dfrac{\beta}{\beta_{\Pi}}\sum_{k=q,g}\!\!\bigg{[}\dfrac{14}{3}\tilde{J}^{(3)}_{k~{}63}+\dfrac{10}{3}\tilde{J}^{(1)}_{k~{}42}+\delta\omega_{k}\Big{(}\dfrac{14}{3}\tilde{L}^{(3)}_{k~{}63}+\dfrac{10}{3}\tilde{L}^{(1)}_{k~{}42}-\dfrac{7}{3}\xi_{k}+\Gamma_{k}\Big{)}\!\bigg{]}, (62)
λΠπ\displaystyle\lambda_{\Pi\pi} =β3βπk=q,g[7J~k63(3)+2J~k52(2)+δωk(7L~k63(3)+2L~k52(2)2ξk)]cs2,\displaystyle=\dfrac{\beta}{3\beta_{\pi}}\sum_{k=q,g}\bigg{[}7\tilde{J}^{(3)}_{k~{}63}+2\tilde{J}^{(2)}_{k~{}52}+\delta\omega_{k}\Big{(}7\tilde{L}^{(3)}_{k~{}63}+2\tilde{L}^{(2)}_{k~{}52}-2\xi_{k}\Big{)}\bigg{]}-c_{s}^{2}, (63)
δΠΠ\displaystyle\delta_{\Pi\Pi} =ββΠk=q,g[59λ0kδωkλ1k+(δωkβ)λ2k(δωk)2λ3k+δωk(δωkβ)λ4k(δωkβ)2λ5k]cs2,\displaystyle=\frac{\beta}{\beta_{\Pi}}\!\!\sum_{k=q,g}\!\!\bigg{[}\!\!-\!\frac{5}{9}\lambda_{0k}\!-\!\delta\omega_{k}\lambda_{1k}\!+\!\!\left(\!\frac{\partial\delta\omega_{k}}{\partial\beta}\!\right)\!\lambda_{2k}\!-\!(\delta\omega_{k})^{2}\lambda_{3k}\!+\!\delta\omega_{k}\!\left(\!\frac{\partial\delta\omega_{k}}{\partial\beta}\!\right)\!\!\lambda_{4k}\!-\!\left(\!\frac{\partial\delta\omega_{k}}{\partial\beta}\!\right)^{2}\!\!\!\lambda_{5k}\!\bigg{]}\!-\!c_{s}^{2}, (64)

where,

ξk\displaystyle\xi_{k} =J~k42(2)+δωkL~k42(2),\displaystyle=~{}\tilde{J}^{(2)}_{k~{}42}+\delta\omega_{k}\tilde{L}^{(2)}_{k~{}42}, (65)
Γk\displaystyle\Gamma_{k} =J~k21(0)βM~k42(1)+δωk(L~k21(0)J~k21(1)βN~k42(1)),\displaystyle=~{}\tilde{J}^{(0)}_{k~{}21}-\beta\tilde{M}^{(1)}_{k~{}42}+\delta\omega_{k}\Big{(}\tilde{L}^{(0)}_{k~{}21}-\tilde{J}^{(1)}_{k~{}21}-\beta\tilde{N}^{(1)}_{k~{}42}\Big{)}, (66)
λ1k\displaystyle\lambda_{1k} =(83J~k21(0)βM~k31(0))cs2+259J~k42(2)53J~k31(1)53βM~k42(1),\displaystyle=~{}\Big{(}\dfrac{8}{3}\tilde{J}^{(0)}_{k~{}21}-\beta\tilde{M}^{(0)}_{k~{}31}\Big{)}c_{s}^{2}+\dfrac{25}{9}\tilde{J}^{(2)}_{k~{}42}-\dfrac{5}{3}\tilde{J}^{(1)}_{k~{}31}-\dfrac{5}{3}\beta\tilde{M}^{(1)}_{k~{}42}, (67)
λ2k\displaystyle\lambda_{2k} =53(J~k31(1)+J~k42(2)+βM~k42(1)L~k42(1))βcs2+M~k31(0)β2(cs2)2,\displaystyle=~{}\dfrac{5}{3}\Big{(}\tilde{J}^{(1)}_{k~{}31}+\tilde{J}^{(2)}_{k~{}42}+\beta\tilde{M}^{(1)}_{k~{}42}-\tilde{L}^{(1)}_{k~{}42}\Big{)}\beta c_{s}^{2}+\tilde{M}^{(0)}_{k~{}31}\beta^{2}(c_{s}^{2})^{2}, (68)
λ3k\displaystyle\lambda_{3k} =53J~k21(1)βM~k21(0)+(83L~k21(0)βN~k31(0))cs2+259L~k42(2)53L~k31(1)53βN~k42(1),\displaystyle=~{}\dfrac{5}{3}\tilde{J}^{(1)}_{k~{}21}-\beta\tilde{M}^{(0)}_{k~{}21}+\Big{(}\dfrac{8}{3}\tilde{L}^{(0)}_{k~{}21}-\beta\tilde{N}^{(0)}_{k~{}31}\Big{)}c_{s}^{2}+\dfrac{25}{9}\tilde{L}^{(2)}_{k~{}42}-\dfrac{5}{3}\tilde{L}^{(1)}_{k~{}31}-\dfrac{5}{3}\beta\tilde{N}^{(1)}_{k~{}42}, (69)
λ4k\displaystyle\lambda_{4k} =(13J~k21(1)+2βM~k21(0))βcs2+N~k31(0)β2(cs2)2+53(35L~k21(0)+L~k31(1)+L~k42(2)+βN~k42(1))βcs2,\displaystyle=~{}\Big{(}\dfrac{1}{3}\tilde{J}^{(1)}_{k~{}21}+2\beta\tilde{M}^{(0)}_{k~{}21}\Big{)}\beta c_{s}^{2}+\tilde{N}^{(0)}_{k~{}31}\beta^{2}(c_{s}^{2})^{2}+\dfrac{5}{3}\Big{(}\dfrac{3}{5}\tilde{L}^{(0)}_{k~{}21}+\tilde{L}^{(1)}_{k~{}31}+\tilde{L}^{(2)}_{k~{}42}+\beta\tilde{N}^{(1)}_{k~{}42}\Big{)}\beta c_{s}^{2}, (70)
λ5k\displaystyle\lambda_{5k} =(J~k21(1)βM~k21(0)+β1L~k31(0))β2(cs2)2.\displaystyle=~{}\Big{(}\tilde{J}^{(1)}_{k~{}21}-\beta\tilde{M}^{(0)}_{k~{}21}+\beta^{-1}\tilde{L}^{(0)}_{k~{}31}\Big{)}\beta^{2}(c_{s}^{2})^{2}. (71)

Thermodynamic integrals appearing in the second order transport coefficients are as shown below

J~knm(r)\displaystyle\tilde{J}^{(r)}_{k~{}nm} =gk2π2(1)m(2m+1)!!0dp~k(u.p~k)n2mr1(p~k)2m+2fk0f¯k0,\displaystyle=\dfrac{g_{k}}{2\pi^{2}}\frac{(-1)^{m}}{(2m+1)!!}\int_{0}^{\infty}{d\mid\vec{\tilde{p}}_{k}\mid}~{}{\big{(}u.\tilde{p}_{k}\big{)}^{n-2m-r-1}}\big{(}\mid\vec{\tilde{p}}_{k}\mid\big{)}^{2m+2}f^{0}_{k}\bar{f}^{0}_{k}, (72)
L~knm(r)\displaystyle\tilde{L}^{(r)}_{k~{}nm} =gk2π2(1)m(2m+1)!!0dp~k(u.p~k)n2mr1Ek(p~k)2m+2fk0f¯k0.\displaystyle=\dfrac{g_{k}}{2\pi^{2}}\frac{(-1)^{m}}{(2m+1)!!}\int_{0}^{\infty}{d\mid\vec{\tilde{p}}_{k}\mid}~{}\dfrac{\big{(}u.\tilde{p}_{k}\big{)}^{n-2m-r-1}}{E_{k}}\big{(}\mid\vec{\tilde{p}}_{k}\mid\big{)}^{2m+2}f^{0}_{k}\bar{f}^{0}_{k}. (73)
M~knm(r)=\displaystyle\tilde{M}^{(r)}_{k~{}nm}= gk2π2(1)m(2m+1)!!0dp~k(u.p~k)n2mr1(p~k)2m+2(f¯k0afk0)fk0f¯k0,\displaystyle~{}\dfrac{g_{k}}{2\pi^{2}}\frac{(-1)^{m}}{(2m+1)!!}\int_{0}^{\infty}{d\mid\vec{\tilde{p}}_{k}\mid}~{}{\big{(}u.\tilde{p}_{k}\big{)}^{n-2m-r-1}}\big{(}\mid\vec{\tilde{p}}_{k}\mid\big{)}^{2m+2}(\bar{f}^{0}_{k}-af^{0}_{k})f^{0}_{k}\bar{f}^{0}_{k}, (74)
N~knm(r)=\displaystyle\tilde{N}^{(r)}_{k~{}nm}= gk2π2(1)m(2m+1)!!0dp~k(u.p~k)n2mr1Ek(p~k)2m+2(f¯k0afk0)fk0f¯k0.\displaystyle~{}\dfrac{g_{k}}{2\pi^{2}}\frac{(-1)^{m}}{(2m+1)!!}\int_{0}^{\infty}{d\mid\vec{\tilde{p}}_{k}\mid}~{}\dfrac{{\big{(}u.\tilde{p}_{k}\big{)}^{n-2m-r-1}}}{E_{k}}\big{(}\mid\vec{\tilde{p}}_{k}\mid\big{)}^{2m+2}(\bar{f}^{0}_{k}-af^{0}_{k})f^{0}_{k}\bar{f}^{0}_{k}. (75)

We note that, these thermodynamic integrals can be approximated and expressed in terms of polylogrithmic functions of order nn and argument aa, PolyLog[n,a][n,a] as shown in Ref. Bhadury et al. (2021).

References