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

Langevin dynamics of heavy quarks in a soft-hard factorized approach

Shuang Li [email protected] College of Science, China Three Gorges University, Yichang 443002, China Key Laboratory of Quark and Lepton Physics (MOE), Central China Normal University, Wuhan 430079, China    Fei Sun [email protected] College of Science, China Three Gorges University, Yichang 443002, China    Wei Xie [email protected] College of Science, China Three Gorges University, Yichang 443002, China    Wei Xiong [email protected] College of Science, China Three Gorges University, Yichang 443002, China
Abstract

By utilizing a soft-hard factorized model, which combines a thermal perturbative description of soft scatterings and a perturbative QCD-based calculation for hard collisions, we study the energy and temperature dependence of the heavy quark diffusion coefficients in Langevin dynamics. The adjustable parameters are fixed from the comprehensive model-data comparison. We find that a small value of the spatial diffusion coefficient at transition temperature is preferred by data 2πTDs(Tc)62\pi TD_{s}(T_{c})\simeq 6. With the parameter-optimized model, we are able to describe simultaneously the prompt D0D^{0} RAAR_{\rm AA} and v2v_{\rm 2} data at pT8p_{\rm T}\leq 8 GeV in Pb–Pb collisions at sNN=2.76\sqrt{s_{\rm NN}}=2.76 and sNN=5.02\sqrt{s_{\rm NN}}=5.02 TeV. We also make predictions for non-prompt D0D^{0} meson for future experimental tests down to the low momentum region.

I INTRODUCTION

Ultrarelativistic heavy-ion collisions provide a unique opportunity to create and investigate the properties of strongly interacting matter in extreme conditions of temperature and energy density, where the normal matter turns into a new form of nuclear matter, consisting of deconfined quarks and gluons, namely quark-gluon plasma (QGP Bzdak et al. (2020)). Such collisions allow us to study the properties of the produced hot and dense partonic medium, which are important for our understanding of the properties of the universe in the first few milliseconds and the composition of the inner core of neutron stars E. Shuryak (1980); Braun-Munzinger and Wambach (2009); Braun-Munzinger et al. (2016). Over the past two decades, the measurements with heavy-ion collisions have been carried at the Relativistic Heavy Ion Collider (RHIC) at BNL and the Large Hadron Collider (LHC) at CERN Gyulassy and L. McLerran (2005); E. Shuryak (2005); B. Muller, J. Schukraft, and B. Wyslouch (2012), to search and explore the fundamental properties of QGP, notably its transport coefficients related to the medium interaction of hard probes.

Heavy quarks (HQs), including charm and bottom, provide a unique insight into the microscopic properties of QGP Zhou et al. (2014); Tang et al. (2014); Andronic et al. (2016); X. Dong and V. Greco (2019); Dong et al. (2019); Zhao et al. (2020). Due to large mass, they are mainly produced in initial hard scatterings and then traverse the QGP and experience elastic and inelastic scatterings with its thermalized constituents Rapp and van Hees (2010); F. Prino and R. Rapp (2016). The transport properties of HQ inside QGP are encoded in the HQ transport coefficients, which are expected to affect the distributions of the corresponding open heavy-flavor hadrons. The resulting experimental observables like the nuclear modification factor RAAR_{\rm AA} and elliptic anisotropy v2v_{\rm 2} of various DD and BB-mesons are therefore sensitive to the HQ transport coefficients, in particular their energy and temperature dependence. There now exist an extensive set of such measurements, which allow a data-based extraction of these coefficients. In this work, we make such an attempt by using a soft-hard factorized model (see Sec. III) to calculate the diffusion and drag coefficients relevant for heavy quark Langevin dynamics.

A particularly important feature of the QGP transport coefficients is their momentum and temperature dependence, especially how they change within the temperature region accessed by the RHIC and LHC experiments. For instance the normalized jet transport coefficient q^/T3\hat{q}/T^{3} was predicted to present a rapidly increasing behavior with decreasing temperature and develop a near-TcT_{c} peak structure J. F. Liao and E. Shuryak (2009). The subsequent studies  J. C. Xu, J. F. Liao and M. Gyulassy (2015, 2016); JET Collaboration (2014); S. Z. Shi, J. F. Liao, and M. Gyulassy (2018, 2019); Ramamurti and Shuryak (2018) seem to confirm this scenario. Another important transport property, shear viscosity over entropy density ratio η/s\eta/s, also presents a visible TT-dependence with a considerable increase above TcT_{c} J. E. Bernhard, J. S. Moreland and S. A. Bass (2019). Concerning the HQ diffusion and drag coefficients, there are also indications of nontrivial temperature dependence from both the phenomenological extractions S. S. Cao, G. Y. Qin, and S. A. Bass (2015); Cao et al. (2016); M. Greif, F. Senzel, H. Kremer, K. Zhou, C. Greiner, and Z. Xu (2017); S. Li, C. W. Wang, X. B. Yuan, and S. Q. Feng (2018); S. Li and C. W. Wang (2018); Prado et al. (2020); Wang et al. (2020) and theoretical calculations M. He, R. J. Fries, and R. Rapp (2013); T. Song, H. Berrehrah, D. Cabrera, W. Cassing, and E. Bratkovskaya (2016); W. M. Alberico, A. Beraudo, A. De Pace, A. Molinari, M. Monteno, M. Nardi, and F. Prino (2011); O. Fochler, Z. Xu, and C. Greiner (2010); S. Z. Shi, J. F. Liao, and M. Gyulassy (2018, 2019). The difference among the derived hybrid models is mainly induced by the treatment of the scale of QCD strong coupling, hadronization and non-perturbative effects Gossiaux (2019); Cao (2021). See Ref. Beraudo et al. (2018); Xu et al. (2019); Cao et al. (2019) for the recent comparisons.

The paper is organized as follows. In Sec. II we introduce the general setup of the employed Langevin dynamics. Section III is dedicated to the detailed calculation of the heavy quark transport coefficients with the factorization model. In Sec. IV we show systematic comparisons between modeling results and data and optimize the model parameters based on global χ2\chi^{2} analysis. With the parameter-optimized model, the energy and temperature dependence of the heavy quark transport coefficients are presented in Sec. V, as well as the comparisons with data and other theoretical calculations. Section VI contains the summary and discussion.

II Langevin dynamics

The classicial Langevin Transport Equation (LTE) of a single HQ reads Rapp and van Hees (2010)

dxi=piEdt\displaystyle dx^{i}=\frac{p^{i}}{E}dt (1a)
dpi=ηDpidt+Cikρkdt.\displaystyle dp^{i}=-\eta_{D}p^{i}dt+C^{ik}\rho^{k}\sqrt{dt}. (1b)

with i,k=1,2,3i,k=1,2,3. The first term on the right hand side of Eq. 1b represents the deterministic drag force, Fdragi=ηDpiF^{i}_{drag}=-\eta_{D}p^{i}, which is given by the drag coefficient ηD(E,T)\eta_{D}(E,T) with the HQ energy E=p 2+mQ2E=\sqrt{\vec{p}^{\;2}+m^{2}_{Q}} and the underlying medium temperature TT. The second term denotes the stochastic thermal force, Fthermali=Cikρk/dtF^{i}_{thermal}=C^{ik}\rho^{k}/\sqrt{dt}, which is described by the momentum argument of the covariance matrix CikC^{ik},

CikκT(δikpipkp 2)+κLpipkp 2,C^{ik}\equiv\sqrt{\kappa_{T}}(\delta^{ik}-\frac{p^{i}p^{k}}{\vec{p}^{\;2}})+\sqrt{\kappa_{L}}\frac{p^{i}p^{k}}{\vec{p}^{\;2}}, (2)

together with a Gaussian-normal distributed random variable ρk\rho^{k}, resulting in the uncorrelated random momentum kicks between two different time scales

<Fthermali(t)Fthermalj(t)>ρ\displaystyle<\vec{F}^{i}_{thermal}(t)\cdot\vec{F}^{j}_{thermal}(t^{\prime})>_{\rho} (3)
=CikCjkδ(tt)\displaystyle=C^{ik}C^{jk}\delta(t-t^{\prime})
=(2)[κT(δijpipjp 2)+κLpipjp 2]δ(tt).\displaystyle\stackrel{{\scriptstyle(\ref{eq:LTEtensor})}}{{=}}\bigr{[}\kappa_{T}(\delta^{ij}-\frac{p^{i}p^{j}}{\vec{p}^{\;2}})+\kappa_{L}\frac{p^{i}p^{j}}{\vec{p}^{\;2}}\bigr{]}\delta(t-t^{\prime}).

κT\kappa_{T} and κL\kappa_{L} are the transverse and longitudinal momentum diffusion coefficients, respectively, which describe the momentum fluctuations in the direction that perpendicular (i.e. transverse) and parallel (i.e. longitudinal) to the propagation. Considering the Einstein relationship, which enforces the drag coefficient starting from the momentum diffusion coefficients as S. Li and J. F. Liao (2020)

ηD=\displaystyle\eta_{D}= κL2TE+(ξ1)κLp 2+d12p 2[ξ(κT+κL)2\displaystyle\frac{\kappa_{L}}{2TE}+(\xi-1)\frac{\partial\kappa_{L}}{\partial\vec{p}^{\;2}}+\frac{d-1}{2\vec{p}^{\;2}}\bigr{[}\xi(\sqrt{\kappa_{T}}+\sqrt{\kappa_{L}})^{2} (4)
(3ξ1)κT(ξ+1)κL].\displaystyle-(3\xi-1)\kappa_{T}-(\xi+1)\kappa_{L}\bigr{]}.

The parameter ξ\xi denotes the discretization scheme of the stochastic integral, which typically takes the values ξ=0,0.5,1\xi=0,0.5,1, representing the pre-point Ito, the mid-point Stratonovic, and the post-point discretization schemes, respectively; d=3d=3 indicates the spatial dimension. In the framework of LTE the HQ-medium interactions are conveniently encoded into three transport coefficients, i.e. ηD\eta_{D}, κT\kappa_{T} and κL\kappa_{L} (Eq. 4). All the problems are therefore reduced to the evaluation of κT/L(E,T)\kappa_{T/L}(E,T), which will be mainly discussed in this work.

Finally, we introduce a few detailed setups of the numerical implementation. The space-time evolution of the temperature field and the velocity field are needed to solve LTE (Eq. 1a and 1b). Following our previous analysis S. Li, C. W. Wang, X. B. Yuan, and S. Q. Feng (2018); S. Li, C. W. Wang, R. Z. Wan, and J. F. Liao (2019), they are obtained in a 3+1 dimensional viscous hydrodynamic calculation I. Karpenko, P. Huoviven and M. Bleicher (2014), with the local thermalization started at τ0=0.6fm/c\tau_{0}=0.6~{}{\rm fm}/{\it c}, the shear viscosity η/s=1/(4π){\eta/s=1/(4\pi)} and the critical temperature Tc=165MeVT_{c}=165~{}{\rm MeV} (see details in Ref. I. Karpenko, P. Huoviven and M. Bleicher (2014)). When the medium temperature drops below TcT_{c}, heavy quark will hadronize into the heavy-flavor hadrons via a fragmentation-coalescence approach. The Braaten-like fragmentation functions are employed for both charm and bottom quarks E. Braaten, K. Cheung, and T. C. Yuan (1993); M. Cacciari, P. Nason and R. Vogt (2005). An instantaneous approach is utilized to characterize the coalescence process for the formation of heavy-flavor mesons from the heavy and (anti-)light quark pairs. The relevant coalescence probability is quantified by the overlap integral of the Winger functions for the meson and partons, which are defined through a harmonic oscillator and the Gaussian wave-function C. B. Dover, U. Heinz, E. Schnedermann, and J. Zimányi (1991), respectively. See Ref. S. Li and J. F. Liao (2020) for more details.

III Momentum diffusion coefficients in a soft-hard factorized approach

When propagating throughout the QGP, the HQ scattering off the gluons and (anti-)partons of the thermal deconfined medium, can be characterized as the two-body elementary processes,

Q(p1)+i(p2)Q(p3)+i(p4),Q~{}(p_{1})+i~{}(p_{2})\rightarrow Q~{}(p_{3})+i~{}(p_{4}), (5)

with p1=(E1,p1)p_{1}=(E_{1},\vec{p}_{1}) and p2p_{2} are the four-momentum of the injected HQ (QQ) and the incident medium partons i=q,gi=q,g, respectively, while p3p_{3} and p4p_{4} are for the ones after scattering. Note that the medium partons are massless (m2=m40m_{2}=m_{4}\sim 0) in particular comparing with the massive HQ (m1=m3=mQm_{1}=m_{3}=m_{Q} in a few times GeV\rm GeV). The corresponding four-momentum transfer is (ω,q)=(ω,qT,qL)(\omega,~{}\vec{q}^{\;})=(\omega,~{}\vec{q}_{T},~{}q_{L}). The Mandelstam invariants read

s(p1+p2)2\displaystyle s\equiv(p_{1}+p_{2})^{2} (6)
t(p1p3)2=ω2q2\displaystyle t\equiv(p_{1}-p_{3})^{2}=\omega^{2}-q^{2}
u(p1p4)2\displaystyle u\equiv(p_{1}-p_{4})^{2}

with q|q|E1q\equiv|\vec{q}\;|\ll E_{1} for small momentum exchange. The transverse and longitudinal momentum diffusion coefficients can be determined by weighting the differential interaction rate with the squared transverse and longitudinal momentum transfer, respectively. It yields

κT\displaystyle\kappa_{T} =12dΓqT 2=12dΓ[ω2t(2ωE1t2|p1|)2]\displaystyle=\frac{1}{2}\int d{\Gamma}\;\vec{q}_{T}^{\;2}=\frac{1}{2}\int d{\Gamma}\bigr{[}\omega^{2}-t-\bigr{(}\frac{2\omega E_{1}-t}{2|\vec{p}_{1}|}\bigr{)}^{2}\bigr{]} (7)

and

κL\displaystyle\kappa_{L} =𝑑ΓqL2=14p1 2𝑑Γ(2ωE1t)2.\displaystyle=\int d{\Gamma}\;q_{L}^{2}=\frac{1}{4\vec{p}_{1}^{\;2}}\int d{\Gamma}\;(2\omega E_{1}-t)^{2}. (8)

As the momentum transfer vanishes (|t|0|t|\rightarrow 0), the gluon propagator in the tt-channel of the elastic process causes an infrared divergence in the squared amplitude |2¯|tchannel1/t2\overline{|\mathcal{M}^{2}}|_{t-channel}\propto 1/t^{2}, which is usually regulated by a Debye screening mass, i.e. ttλmD2t\rightarrow t-\lambda m^{2}_{D} with an adjustable parameter λ\lambda B. Svetitsky (1988); P. B. Gossiaux and J. Aichelin (2008). Alternatively, it can be overcome by utilizing a soft-hard factorized approach S. Peigne´\acute{\rm e} and A. Peshier (2008a, b), which starts with the assumptions that the medium is thermal and weakly coupled, and then the interactions between the heavy quarks and the medium can be computed in thermal perturbation theory. Finally, this approach allows to decompose the soft HQ-medium interactions with t>tt>t^{\ast}, from the hard ones with t<tt<t^{\ast}. For soft collisions the gluon propagator should be replaced by the hard-thermal loop (HTL) propagator E. Braaten and T. C. Yuan (1991); J. P. Blaizot and E. Iancu (2002), while for hard collisions the hard gluon exchange is considered and the Born approximation is appropriate. Therefore, the final results of κT/L\kappa_{T/L} include the contributions from both soft and hard components.

As discussed in the QED case S. Peigne´\acute{\rm e} and A. Peshier (2008a), μ+γμ+γ\mu+\gamma\rightarrow\mu+\gamma, the complete calculation for the energy loss of the energic incident heavy-fermion is independent of the intermediate scale tt^{\ast} in high energy limit. However, in the QCD case, there is the complication that the challenge of the validity of the HTL scenario, mD2T2m_{D}^{2}\ll T^{2} S. Peigne´\acute{\rm e} and A. Peshier (2008b), due to the temperatures reached at RHIC and LHC energies. Consequently, in the QCD case, the soft-hard approach is in fact not independent of the intermediate scale tt^{\ast} P. B. Gossiaux and J. Aichelin (2008); W. M. Alberico, A. Beraudo, A. De Pace, A. Molinari, M. Monteno, M. Nardi, and F. Prino (2011). In this analysis we have checked that the calculations for κT/L\kappa_{T/L} are not sensitive to the choice of the artificial cutoff tmD2t^{\ast}\sim m_{D}^{2}.

In the next parts of this section, we will focus on the energy and temperature dependence of the interaction rate Γ{\Gamma} at leading order in gg for the elastic process, as well as the momentum diffusion coefficients (Eq. 7 and 8) in soft and hard collisions, respectively.

III.1 κT/L\kappa_{T/L} in soft region t<t<0t^{\ast}<t<0

In soft collisions the exchanged four-momentum is soft, tgT\sqrt{-t}\sim gT (λmfp1/g2T\lambda_{mfp}\sim 1/g^{2}T J. Ghiglieri, G. Moore, and D. Teaney (2016)), and the tt-channel long-wavelength gluons are screened by the mediums, thus, they feel the presence of the medium and require the resummation. Here we just show the final results, and the details are relegated to A. The transverse and longitudinal momentum diffusion coefficients can be expressed as

κT(E1,T)=\displaystyle\kappa_{T}(E_{1},T)= CFg216π2v13t0𝑑t(t)3/20v1𝑑xv12x2(1x2)5/2\displaystyle\frac{C_{F}g^{2}}{16\pi^{2}v_{1}^{3}}\int^{0}_{t^{\ast}}dt\;(-t)^{3/2}\int_{0}^{v_{1}}dx\frac{v_{1}^{2}-x^{2}}{(1-x^{2})^{5/2}} (9)
[ρL(t,x)+(v12x2)ρT(t,x)]coth(x2Tt1x2)\displaystyle\bigr{[}\rho_{L}(t,x)+(v_{1}^{2}-x^{2})\rho_{T}(t,x)\bigr{]}coth\bigr{(}\frac{x}{2T}\sqrt{\frac{-t}{1-x^{2}}}\bigr{)}

and

κL(E1,T)=\displaystyle\kappa_{L}(E_{1},T)= CFg28π2v13t0𝑑t(t)3/20v1𝑑xx2(1x2)5/2\displaystyle\frac{C_{F}g^{2}}{8\pi^{2}v_{1}^{3}}\int^{0}_{t^{\ast}}dt\;(-t)^{3/2}\int_{0}^{v_{1}}dx\frac{x^{2}}{(1-x^{2})^{5/2}} (10)
[ρL(t,x)+(v12x2)ρT(t,x)]coth(x2Tt1x2),\displaystyle\bigr{[}\rho_{L}(t,x)+(v_{1}^{2}-x^{2})\rho_{T}(t,x)\bigr{]}coth\bigr{(}\frac{x}{2T}\sqrt{\frac{-t}{1-x^{2}}}\bigr{)},

respectively, with the HQ velocity v1=|p|/E1v_{1}=|\vec{p}|/E_{1} and the transverse and longitudinal parts of the HTL gluon spectral functions444 The spectral function involves only the low frequency excitations, namely the Landau cut, while the quasiparticle excitations is irrelevant in this regime. See Eq.25 for more details. are given by

ρT(t,x)=\displaystyle\rho_{T}(t,x)= πmD22x(1x2){[t+mD22x2(1+1x22xln1+x1x)]2\displaystyle\frac{\pi m_{D}^{2}}{2}x(1-x^{2})\biggr{\{}\bigr{[}-t+\frac{m_{D}^{2}}{2}x^{2}(1+\frac{1-x^{2}}{2x}ln\frac{1+x}{1-x})\bigr{]}^{2} (11)
+[πmD24x(1x2)]2}1\displaystyle+\bigr{[}\frac{\pi m_{D}^{2}}{4}x(1-x^{2})\bigr{]}^{2}\biggr{\}}^{-1}

and

ρL(t,x)=\displaystyle\rho_{L}(t,x)= πmD2x{[t1x2+mD2(1x2ln1+x1x)]2\displaystyle\pi m_{D}^{2}x\biggr{\{}\bigr{[}\frac{-t}{1-x^{2}}+m_{D}^{2}(1-\frac{x}{2}ln\frac{1+x}{1-x})\bigr{]}^{2} (12)
+(πmD22x)2}1.\displaystyle+\bigr{(}\frac{\pi m_{D}^{2}}{2}x\bigr{)}^{2}\biggr{\}}^{-1}.

III.2 κT/L\kappa_{T/L} in hard region tmin<t<tt_{min}<t<t^{\ast}

In hard collisions the exchanged four-momentum is hard, tT\sqrt{-t}\gtrsim T (λmfp1/g4T\lambda_{mfp}\sim 1/g^{4}T J. Ghiglieri, G. Moore, and D. Teaney (2016)), and the pQCD Born approximation is valid in this regime. In analogy with the previous part we give the κT/L\kappa_{T/L} results directly, and the detailed aspects of the calculations can be found in B. The momentum diffusion coefficients reads

κTQi(E1,T)=\displaystyle\kappa^{Qi}_{T}(E_{1},T)= 1256π3|p1|E1|p2|mind|p2|E2n2(E2)\displaystyle\frac{1}{256\pi^{3}|\vec{p}_{1}|E_{1}}\int_{|\vec{p}_{2}|_{min}}^{\infty}d|\vec{p}_{2}|E_{2}n_{2}(E_{2}) (13)
1cosψ|maxd(cosψ)tmintdt1a[m12(D+2b2)8p1 2a4\displaystyle\int_{-1}^{cos\psi|_{max}}d(cos\psi)\int_{t_{min}}^{t^{\ast}}dt\frac{1}{a}\bigr{[}-\frac{m_{1}^{2}(D+2b^{2})}{8\vec{p}_{1}^{\;2}a^{4}}
+E1tb2p1 2a2t(1+t4p1 2)]|2¯|Qi\displaystyle+\frac{E_{1}tb}{2\vec{p}_{1}^{\;2}a^{2}}-t(1+\frac{t}{4\vec{p}_{1}^{\;2}})\bigr{]}\;\overline{|\mathcal{M}^{2}}|^{Qi}

and

κLQi(E1,T)=\displaystyle\kappa^{Qi}_{L}(E_{1},T)= 1256π3|p1|3E1|p2|mind|p2|E2n2(E2)\displaystyle\frac{1}{256\pi^{3}|\vec{p}_{1}|^{3}E_{1}}\int_{|\vec{p}_{2}|_{min}}^{\infty}d|\vec{p}_{2}|E_{2}n_{2}(E_{2}) (14)
1cosψ|maxd(cosψ)tmintdt1a[E12(D+2b2)4a4\displaystyle\int_{-1}^{cos\psi|_{max}}d(cos\psi)\int_{t_{min}}^{t^{\ast}}dt\frac{1}{a}\bigr{[}\frac{E_{1}^{2}(D+2b^{2})}{4a^{4}}
E1tba2+t22]|2¯|Qi.\displaystyle-\frac{E_{1}tb}{a^{2}}+\frac{t^{2}}{2}\bigr{]}\;\overline{|\mathcal{M}^{2}}|^{Qi}.

The integrations limits and the short notations are shown in Eq. 45-51.

III.3 Complete results in soft-hard scenario

Combining the soft and hard contributions to the momentum diffusion coefficients via

κT/L(E,T)\displaystyle\kappa_{T/L}(E,T) =κT/Lsoft(E,T)+κT/Lhard(E,T)\displaystyle=\kappa_{T/L}^{soft}(E,T)+\kappa_{T/L}^{hard}(E,T) (15)
=κT/Lsoft(E,T)+i=q,gκT/LhardQi\displaystyle=\kappa_{T/L}^{soft}(E,T)+\sum_{i=q,g}\kappa_{T/L}^{hard-Qi}

while κT/Lsoft(EE1,T)\kappa_{T/L}^{soft}(E\equiv E_{1},T) is given by Eq. 9 and 10, and κT/LhardQi\kappa_{T/L}^{hard-Qi} is expressed in Eq. 13 and 14 for a given incident medium parton i=q,gi=q,g. Adopting the post-point discretization scheme of the stochastic integral, i.e. ξ=1\xi=1 in Eq. 4, the drag coefficient ηD(E,T)\eta_{D}(E,T) can be obtained by inserting Eq. 15 into Eq. 4.

IV Data-based parameter optimization

Following the strategies utilized in our previous work S. Li and J. F. Liao (2020); S. Li, W. Xiong, and R. Z. Wan (2020), the two key parameters in this study, the intermediate cutoff tt^{\ast} and the scale μ\mu of running coupling (Eq. 31a), are tested within a wide range of possibility and drawn constrains by comparing the relevant charm meson data with model results. We calculate the corresponding final observable y for the desired species of D-meson. Then, a χ2\chi^{2} analysis can be performed by comparing the model predictions with experimental data

χ2=i=1N(𝐲𝐢Data𝐲𝐢Modelσi)2.\displaystyle\chi^{2}=\sum_{i=1}^{N}\biggr{(}\frac{\bf y^{\rm Data}_{i}-y^{\rm Model}_{i}}{\sigma_{i}}\biggr{)}^{2}. (16)

In the above σi\sigma_{i} is the total uncertainty in data points, including the statistic and systematic components which are added in quadrature. n=N1n=N-1 denotes the degree of freedom (d.o.fd.o.f) when there are NN data points used in the comparison. In this study, we use an extensive set of LHC data in the range pT8GeVp_{\rm T}\leq 8~{}\rm{GeV}: D0D^{0}, D+D^{+}, D+D^{\ast+} and Ds+D_{s}^{+} RAAR_{\rm AA} data collected at mid-rapidity (|y|<0.5|y|<0.5) in the most central (010%0-10\%) and semi-central (3050%30-50\%) Pb–Pb collisions at sNN=2.76TeV\sqrt{s_{\rm NN}}=2.76~{}{\rm TeV} ALICE Collaboration (2016a, b) and sNN=5.02TeV\sqrt{s_{\rm NN}}=5.02~{}{\rm TeV} ALICE Collaboration (2018a), as well as the v2v_{\rm 2} data in semi-central (3050%30-50\%) collisions ALICE Collaboration (2014); ALICE Collaboration (2018b); CMS Collaboration (2018a).

We scan a wide range of valus for (tt^{\ast}, μ\mu): 1|t|mD231\leq\frac{|t^{\ast}|}{m_{D}^{2}}\leq 3 and 1μπT31\leq\frac{\mu}{\pi T}\leq 3. A total of 20 different combinations were computed and compared with the experimental data. The obtained results are summarized in Tab. 1. The χ2\chi^{2} values are computed separately for RAAR_{\rm AA} and v2v_{\rm 2} as well as for all data combined. To better visualize the results, we also show them in Fig. 1, with left panels for RAAR_{\rm AA} analysis and right panels for v2v_{2} analysis. In both panels, the y-axis labels the desired parameters, |t|mD2\frac{|t^{\ast}|}{m_{D}^{2}} (upper) and μπT\frac{\mu}{\pi T} (lower), and x-axis labels the χ2/d.o.f\chi^{2}/d.o.f within the selected ranges555χ2/d.o.f\chi^{2}/d.o.f is shown in the range 0.5<χ2/d.o.f<2.50.5<\chi^{2}/d.o.f<2.5 for better visualization.. The different points (filled gry circles) represent the different combinations of parameters (|t|mD2\frac{|t^{\ast}|}{m_{D}^{2}}, μπT\frac{\mu}{\pi T}) in Tab. 1, with the number on top of each point to display the relevant “ModelIDModel~{}ID” for that model. A number of observations can be drawn from the comprehensive model-data comparison. For the RAAR_{\rm AA}, several models achieve χ2/d.o.f1\chi^{2}/d.o.f\simeq 1 with |t|mD21.5\frac{|t^{\ast}|}{m_{D}^{2}}\lesssim 1.5 and widespread values of μ\mu: 1μπT31\leq\frac{\mu}{\pi T}\leq 3. This suggests that RAAR_{\rm AA} appears to be more sensitive to the intermediate cutoff while insensitive to the scale of coupling constant. For the v2v_{\rm 2}, it clearly shows a stronger sensitivity to μ\mu, which seems to give a better description (χ2/d.o.f1.52.0\chi^{2}/d.o.f\simeq 1.5-2.0) of the data with μπT=1\frac{\mu}{\pi T}=1. It is interesting to see that RAAR_{\rm AA} data is more powerful to constrain |t|mD2\frac{|t^{\ast}|}{m_{D}^{2}}, while v2v_{\rm 2} data is more efficient to nail down μπT\frac{\mu}{\pi T}. Taken all together, we can identify a particular model that outperforms others in describing both RAAR_{\rm AA} and v2v_{\rm 2} data simultaneously with χ2/d.o.f=1.3\chi^{2}/d.o.f=1.3. This one will be the parameter-optimized model in this work: |t|=1.5mD2|t^{\ast}|=1.5m_{D}^{2} and μ=πT\mu=\pi T.

𝑀𝑜𝑑𝑒𝑙𝐼𝐷\it Model~{}ID |t|/mD2|t^{\ast}|/m_{D}^{2} μ/πT\mu/\pi T χ2/d.o.f\chi^{2}/d.o.f χ2/d.o.f\chi^{2}/d.o.f Total
(Cutoff) (Scale) (RAAR_{\rm AA}) (v2v_{\rm 2})
1 1.00 1.00 1.02 5.08 1.61
2 1.00 1.50 2.46 6.08 2.98
3 1.00 2.00 1.71 6.11 2.35
4 1.00 3.00 0.83 4.20 1.32
5 1.50 1.00 1.16 2.10 1.30
6 1.50 1.50 3.90 9.82 4.76
7 1.50 2.00 5.38 10.90 6.18
8 1.50 3.00 4.85 9.12 5.47
9 2.00 1.00 2.44 1.54 2.31
10 2.00 1.50 2.99 7.11 3.58
11 2.00 2.00 6.81 10.19 7.30
12 2.00 3.00 8.91 8.19 8.80
13 2.50 1.00 4.04 1.61 3.68
14 2.50 1.50 2.08 6.78 2.77
15 2.50 2.00 6.59 8.51 6.87
16 2.50 3.00 11.80 12.04 11.83
17 3.00 1.00 5.46 1.46 4.88
18 3.00 1.50 1.52 8.05 2.47
19 3.00 2.00 6.27 9.88 6.79
20 3.00 3.00 12.94 15.21 13.27
Table 1: Summary of the adjustable parameters in this work, together with the relevant χ2/d.o.f\chi^{2}/d.o.f obtained for RAAR_{\rm AA} and v2v_{\rm 2}.
Refer to caption
Figure 1: Comparision of χ2/d.o.f\chi^{2}/d.o.f based on the experimental data of RAAR_{\rm AA} (left) and v2v_{\rm 2} (right). The model predictions are calculated by using various combinations of parameter |t|/mD2|t^{\ast}|/m_{D}^{2} (upper) and μ/πT\mu/\pi T (lower), which are represented as χ2/d.o.f\chi^{2}/d.o.f (x-axis) and the desired parameter (y-axis), respectively. See the legend and text for details.

V Results

In this section we will first examine the tt^{\ast} dependence of κT/L\kappa_{T/L} (Eq. 15) for both charm and bottom quark. Then, the relevant energy and temperature dependence of κT/L(E,T)\kappa_{T/L}(E,T) will be discussed with the optimized parameters. For the desired observables we will perform the comparisons with the results from lattice QCD at zero momentum limit, as well as the ones from experimental data in the low to intermediate pTp_{\rm T} region.

V.1 Energy and temperature dependence of the transport coefficients

In Fig. 2, charm (left) and bottom quark (right) κT\kappa_{T} (thin curves) and κL\kappa_{L} (thick curves) are calculated, with mD2|t|2.5mD2m_{D}^{2}\leq{|t^{\ast}|}\leq 2.5m_{D}^{2} and μ=πT\mu=\pi T, at the temperature T=0.40T=0.40 GeV (upper) and the energy E=10.0E=10.0 GeV (lower). κT\kappa_{T} and κL\kappa_{L} are, as expected, identical at zero momentum limit (E=mQE=m_{Q}), while the latter one has a much stronger energy dependence at larger momentum. Furthermore, κT/L(E,T)\kappa_{T/L}(E,T) behave a mild sensitivity to the intermediate cutoff tt^{\ast} W. M. Alberico, A. Beraudo, A. De Pace, A. Molinari, M. Monteno, M. Nardi, and F. Prino (2011). Because the soft-hard approach is strictly speaking valid when the coupling is small mD2T2m_{D}^{2}\ll T^{2} S. Peigne´\acute{\rm e} and A. Peshier (2008b), thus, the above observations support the validity of this approach within the temperature regions even though the coupling is not small.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 2: (Color online) Comparison of the momentum diffusion coefficients κT\kappa_{T} (thin black curves) and κL\kappa_{L} (thick red curves), for charm (left) and bottom quarks (right), displaying separately the results based various testing parameters: mD2|t|2.5mD2m_{D}^{2}\leq{|t^{\ast}|}\leq 2.5m_{D}^{2} and μ=πT\mu=\pi T.

In Fig. 3, charm quark κT\kappa_{T} (left) and κL\kappa_{L} (middle) are evaluated with the optimized parameters, |t|=1.5mD2|t^{\ast}|=1.5m_{D}^{2} and μ=πT\mu=\pi T, including both the soft (dotted blue curves) and hard contributions (dashed black curves), at fixed temperature T=0.40T=0.40 GeV (upper) and at fixed energy E=10.0E=10.0 GeV (lower). It is found that the soft components are significant at low energy/temperature, while they are compatible at larger values. The combined results (solid red curves) are presented as well for comparison. With the post-point scheme (ξ=1\xi=1 in Eq. 15), the drag coefficients (right) behave (1) a nonmonotonic dependence, in particular on the temperature, which is in part due to the improved treatment of the screening in soft collisions, and in part due to the procedure of inferring ηD\eta_{D} from κT/L\kappa_{T/L} F. Prino and R. Rapp (2016); (2) a weak energy dependence at E2mc=3E\gtrsim 2m_{c}=3 GeV. Similar conclusions can be drawn for bottom quark, as shown in Fig. 4.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3: (Color online) Charm quark κT\kappa_{T} (left) and κL\kappa_{L} (middle) are shown at fixed temperature T=0.40T=0.40 GeV (upper) and at fixed energy E=10.0E=10.0 GeV (lower), contributed by the soft (dotted blue curves) and hard collisions (dashed black curves). The combined results (solid red curves) are shown as well for comparison. The derived drag coefficient ηD\eta_{D} (right; Eq. 4) are obtained with the post-point scenario.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 4: Same as Fig. 3 but for bottom quark.

Figure 5 presents the transport coefficient of charm quark, q^=2κT\hat{q}=2\kappa_{T}, at fix momentum p=10p=10 GeV (solid red curve). It can be seen that q^/T3\hat{q}/T^{3} reaches the maximum near the critical temperature, and then followed by a decreasing trend with TT, providing a good description of the light quark transport parameter (black circle points). The results from various phenomenological extractions and theoretical calculations, including a phenomenological fitting analysis with the Langevin-transport with Gluon Radiation (LGR; dotted blue curve S. Li and J. F. Liao (2020)), a LO calculation with a Linearized Boltzmann Diffusion Model (LIDO333LIDO results are shown with only elastic scattering channels.; dashed black curve W. Y. Ke, Y. R. Xu, and S. A. Bass (2018)), a nonperturbative treatment with Quasi-Particle Model (QPM in Catania; dot-dashed green curve F. Scardina, S. K. Das, V. Minissale, S. Plumari, and V. Greco (2017)), a novel confinement with semi-quark-gluon-monopole plasma approach (CUJET3; shadowed red band J. C. Xu, J. F. Liao and M. Gyulassy (2016); S. Z. Shi, J. F. Liao, and M. Gyulassy (2019, 2018)), are displayed as well for comparison. Similar temperature dependence can be observed except the CUJET3 approach, which shows a strong enhancement near TcT_{c} regime.

Refer to caption
Figure 5: (Color online) Transport coefficient, q^/T3(T)\hat{q}/T^{3}(T), of charm quark from the various calculations, including: the soft-hard factorized approach (solid red curve), the LGR model with data optimized parameters (dotted blue curve S. Li and J. F. Liao (2020)), a Bayesian ayalysis from LIDO (dashed black curve W. Y. Ke, Y. R. Xu, and S. A. Bass (2018)), a quasi-particle model from Catania (dot-dashed green curve G. Coci (2018)), CUJET3 (shadowed red band J. C. Xu, J. F. Liao and M. Gyulassy (2016); S. Z. Shi, J. F. Liao, and M. Gyulassy (2019, 2018)) and JET Collaboration (black circle points JET Collaboration (2014)) at p=10p=10 GeV.

The scaled spatial diffusion coefficient describes the low energy interaction strength of HQ in medium G. D. Moore and D. Teaney (2005),

2πTDs=limEmQ2πT2mQηD(E,T),2\pi TD_{s}=\lim_{E\rightarrow m_{Q}}\frac{2\pi T^{2}}{m_{Q}\cdot\eta_{D}(E,T)}, (17)

and it can be calculated by substituting Eq. 4 into Eq. 17. The obtained result for charm quark (mc=1.5m_{c}=1.5 GeV) is displayed as the solid red curve in Fig. 6. It is found that a relatively strong increase of 2πTDs(T)2\pi TD_{s}(T) from crossover temperature TcT_{c} toward high temperature. Meanwhile, the v2v_{\rm 2} data prefers a small value of 2πTDs2\pi TD_{s} near TcT_{c}, 2πTDs(Tc)362\pi TD_{s}(T_{c})\simeq 3-6444This range is estimated from the testing models as displayed in the right panels of Fig. 1., which is close to the lattice QCD calculations D. Banerjee, S. Datta, R. Gavai, and P. Majumdar (2012); H. T. Ding, A. Francis, O. Kaczmarek, F. Karsch, H. Satz, and W. Soeldner (2012); O. Kaczmarek (2014); N. Brambilla, V. Leino, P. Petreczky, and A. Vairo (2020); L. Altenkort, A. M. Eller, O. Kaczmarek, L. Mazur, G. D.Moore, and H. T. Shu (2020). The relevant results from other theoretical analyses, such as LGR S. Li and J. F. Liao (2020), LIDO W. Y. Ke, Y. R. Xu, and S. A. Bass (2018), Catania F. Scardina, S. K. Das, V. Minissale, S. Plumari, and V. Greco (2017) and CUJET3555CUJET3 results are obtained by performing the energy interpolation down to E=mQE=m_{Q}. J. C. Xu, J. F. Liao and M. Gyulassy (2016); S. Z. Shi, J. F. Liao, and M. Gyulassy (2019, 2018), show a similar trend but with much weaker temperature dependence.

Refer to caption
Figure 6: (Color online) Spatial diffusion constant 2πTDs(T)2\pi TD_{s}(T) of charm quark from the sof-hard scenario (solid red curve), LGR (dotted blue curve S. Li and J. F. Liao (2020), LIDO (dashed black curve W. Y. Ke, Y. R. Xu, and S. A. Bass (2018)), Catania (dot-dashed green curve F. Scardina, S. K. Das, V. Minissale, S. Plumari, and V. Greco (2017)), CUJET3 (shadowed red band J. C. Xu, J. F. Liao and M. Gyulassy (2016); S. Z. Shi, J. F. Liao, and M. Gyulassy (2019, 2018)) and lattice QCD calculations (black circle D. Banerjee, S. Datta, R. Gavai, and P. Majumdar (2012), blue triangle H. T. Ding, A. Francis, O. Kaczmarek, F. Karsch, H. Satz, and W. Soeldner (2012), pink square O. Kaczmarek (2014), red inverted triangle N. Brambilla, V. Leino, P. Petreczky, and A. Vairo (2020) and green plus L. Altenkort, A. M. Eller, O. Kaczmarek, L. Mazur, G. D.Moore, and H. T. Shu (2020)). The result for D-meson (long dashed pink curve L. Tolos and J. M. Torres-Rincon (2013)) in the hadronic phase is shown for comparison.

V.2 Comparison with experimental data: RAAR_{\rm AA} and v2v_{\rm 2}

Figure 7 shows the RAAR_{\rm AA} of D0D^{0} (a), D+D^{+} (b), D+D^{\ast+} (c) and Ds+D_{s}^{+} (d) in the most central (010%0-10\%) Pb–Pb collisions at sNN=2.76\sqrt{s_{\rm NN}}=2.76 TeV, respectively. The calculations are done with FONLL initial charm quark spectra and EPS09 NLO parametrization for the nPDF in Pb S. Li, C. W. Wang, X. B. Yuan, and S. Q. Feng (2018), and the pink band reflects the theoretical uncertainties coming from these inputs. It can be seen that, within the experimental uncertainties, the model calculations provide a very good description of the measured pTp_{\rm T}-dependent RAAR_{\rm AA} data for various charm mesons. Concerning the results in Pb–Pb collisions at sNN=5.02\sqrt{s_{\rm NN}}=5.02 TeV, as shown in Fig. 8, a good agreement is found between the model and the measurement at pT6GeV/cp_{\rm T}\lesssim 6~{}{\rm GeV/\it{c}}, while a slightly larger discrepancy observed at larger pTp_{\rm T}.

Refer to caption
Figure 7: (Color online) Comparison between experimental data (red box ALICE Collaboration (2016a, b)) and soft-hard factorized model calculations (solid black curve with pink uncertainty band) for the nuclear modification factor RAAR_{\rm AA}, of D0D^{0} (a), D+D^{+} (b), D+D^{\ast+} (c) and Ds+D_{s}^{+} (d) at mid-rapidity (|y|<0.5|y|<0.5) in central (010%0-10\%) Pb–Pb collisions at sNN=2.76\sqrt{s_{\rm NN}}=2.76 TeV.
Refer to caption
Figure 8: Same as Fig. 7 but for Pb–Pb collisions at sNN=5.02\sqrt{s_{\rm NN}}=5.02 TeV. The data (solid ALICE Collaboration (2018a), open Vertesi (2019)) are shown for comparison.

Figure 9 presents the elliptic flow coefficient v2v_{\rm 2} of non-strange D-meson (averaged D0D^{0}, D+D^{+}, and D+D^{\ast+}) in semi-central (3050%30-50\%) Pb–Pb collisions at sNN=2.76\sqrt{s_{\rm NN}}=2.76 TeV (a) and sNN=5.02\sqrt{s_{\rm NN}}=5.02 TeV (b). Within the uncertainties of the experimental data, our model calculations describe well the anisotropy of the transverse momentum distribution of the non-strange D-meson. The sizable v2v_{\rm 2} of these charm mesons, in particular at intermediate pT35p_{\rm T}\simeq 3-5 GeV, suggests that charm quarks actively participate in the collective expansion of the fireball.

Refer to caption
Refer to caption
Figure 9: (Color online) Comparison between experimental data (red ALICE Collaboration (2014), black ALICE Collaboration (2018b) and blue boxes CMS Collaboration (2018b)) and model calculations (solid black curve with pink uncertainty band) for the elliptic flow v2v_{\rm 2} of non-strange D-meson at mid-rapidity (|y|<0.5|y|<0.5) in semi-central (3050%30-50\%) Pb–Pb collisions at sNN=2.76\sqrt{s_{\rm NN}}=2.76 TeV (a) and sNN=5.02\sqrt{s_{\rm NN}}=5.02 TeV (b).

In Fig. 10, the pTp_{\rm T}-differential RAAR_{\rm AA} of non-prompt D0D^{0} mesons are predicted with the parameter-optimized model, and shown as a function of pTp_{\rm T} in central (010%0-10\%) Pb–Pb collisions at sNN=5.02\sqrt{s_{\rm NN}}=5.02 TeV. Comparing with prompt D0D^{0} RAAR_{\rm AA} (solid curve), at moderate pTp_{\rm T} (pT57p_{\rm T}\sim 5-7), a less suppression behavior is found for D0D^{0} from BB-hadron decays (dashed curve), reflecting a weaker in-medium energy loss effect of bottom quark (mb=4.75m_{b}=4.75 GeV), which has larger mass with respect to that of charm quark (mc=1.5m_{c}=1.5 GeV). We note that the future measurements performed for the non-prompt D0D^{0} RAAR_{\rm AA} and v2v_{\rm 2}, are powerful in nailing down the varying ranges of the model parameters (see Sec. IV), which will largely improve and extend the current understanding of the in-medium effects.

Refer to caption
Figure 10: Comparison of non-prompt (dashed blue curve) and prompt D0D^{0} (solid black curve) RAAR_{\rm AA} calculations as a function of pTp_{\rm T} with the measured values (point ALICE Collaboration (2018a)), in the central 010%0-10\% Pb–Pb collisions at sNN=5.02\sqrt{s_{\rm NN}}=5.02 TeV. See legend and text for details.

VI Summary

In this work we have used a soft-hard factorized model to investigate the heavy quark momentum diffusion coefficients κT/L\kappa_{T/L} in the quark-gluon plasma in a data-driven approach. In particular we’ve examined the validity of this scenario by systematically scanning a wide range of possibilities. The global χ2\chi^{2} analysis using an extensive set of LHC data on charm meson RAAR_{\rm AA} and v2v_{\rm 2} has allowed us to constrain the preferred range of the two parameters: soft-hard intermediate cutoff |t|=1.5mD2|t^{\ast}|=1.5m_{D}^{2} and the scale of QCD coupling constant μ=πT\mu=\pi T. It is found that κT/L\kappa_{T/L} have a mild sensitivity to tt^{\ast}, supporting the validity of the soft-hard approach when the coupling is not small. With this factorization model, we have calculated the transport coefficient q^\hat{q}, drag coefficient ηD\eta_{D}, spatial diffusion coefficient 2πTDs2\pi TD_{s}, and then compared with other theoretical calculations and phenomenological extractions. Our analysis suggests that a small value 2πTDs62\pi TD_{s}\simeq 6 appears to be much preferred near TcT_{c}. Finally we’ve demonstrated a simultaneous description of charm meson RAAR_{\rm AA} and v2v_{\rm 2} observables in the range pT8p_{\rm T}\leq 8 GeV. We’ve further made predictions for bottom meson observables in the same model.

We end with discussions on a few important caveats in the present study that call for future studies:

  • \bullet

    In the framework of Langevin approach, the heavy flavor dynamics is encoded into three coefficient, κT\kappa_{T}, κL\kappa_{L} and ηD\eta_{D}, satisfying Einstein’s relationship. It means that two of them are independent while the third one can be obtained accordingly. The final results therefore depend on the arbitrary choice of which of the two coefficients are calculated with the employed model F. Prino and R. Rapp (2016); Xu et al. (2019). For consistency, we calculate independently the momentum diffusion coefficients κT/L\kappa_{T/L} with the factorization approach, and obtain the drag coefficient via ηD=ηD(κT,κL)\eta_{D}=\eta_{D}(\kappa_{T},\kappa_{L}) (see Eq. 4). A systematic study among the different options may help remedy this situation.

  • \bullet

    According to the present modeling, the resulting spatial diffusion coefficient exhibits a relatively strong increase of temperature comparing with lattice QCD calculations, in particular at large TT, as shown in Tab. 2. It corresponds to a larger relaxation time and thus weaker HQ-medium coupling strength. This comparison can be improved by (1) determining the key parameters based on the upcoming measurements on high precision observables (such as RAAR_{\rm AA}, v2v_{\rm 2} and v3v_{\rm 3}) of both DD and BB-mesons at low pTp_{\rm T}; (2) replacing the current χ2\chi^{2} analysis with a state-of-the-art deep learning technology.

    T=TcT=T_{c} T=2TcT=2T_{c} T=3TcT=3T_{c}
    This work \sim6 \sim22 \sim31
    LQCD (median value) D. Banerjee, S. Datta, R. Gavai, and P. Majumdar (2012); N. Brambilla, V. Leino, P. Petreczky, and A. Vairo (2020) \sim5 \sim10 \sim13
    Table 2: Summary of the different models for 2πTDs2\pi TD_{s} at desired temperature values.
  • \bullet

    It is realized Cao et al. (2020) that the heavy quark hadrochemistry, the abundance of various heavy flavor hadrons, provides special sensitivity to the heavy-light coalescence mechanism and thus plays an important role to understand the observables like the baryon production and the baryon-to-meson ratio. A systematic comparison including the charmed baryons over a broad momentum region is therefore crucial for a better constraining of the model parameters, as well as a better extraction of the heavy quark transport coefficients in a model-to-data approach.

  • \bullet

    The elastic scattering (222\rightarrow 2) processes between heavy quark and QGP constituents are dominated for heavy quark with low to moderate transverse momentum S. Li, C. W. Wang, R. Z. Wan, and J. F. Liao (2019). Thus, here we consider only the elastic energy loss mechanisms to study the observables at pT8p_{\rm T}\leq 8 GeV. The missing radiative (232\leftrightarrow 3) effects may help to reduce the discrepancy with RAAR_{\rm AA} data in the vicinity of pT=8p_{\rm T}=8 GeV, as mentioned above (see Fig. 8). With the soft-hard factorized approach, it would be interesting to include both elastic and radiative contributions in a simultaneous best fit to data in the whole pTp_{\rm T} region. We also plan to explore this idea in the future.

Acknowledgements.
The authors are grateful to Andrea Beraudo and Jinfeng Liao for helpful discussions and communications. We also thank Weiyao Ke and Gabriele Coci for providing the inputs as shown in Fig. 5 and 6. This work is supported by the Hubei Provincial Natural Science Foundation under Grant No.2020CFB163, the National Science Foundation of China (NSFC) under Grant Nos.12005114, 11847014 and 11875178, and the Key Laboratory of Quark and Lepton Physics Contracts Nos.QLPL2018P01 and QLPL201905.

Appendix A Derivation of the interaction rate and momentum diffusion coefficients in soft collisions

In the QED case, μ+γμ+γ\mu+\gamma\rightarrow\mu+\gamma, one can calculate relevant interaction rate with small momentum transfer by using the imaginary part of the muon self-energy Σ(p1)\Sigma(p_{1}) H. A. Weldon (1983)

Γ(E1,T)\displaystyle{\Gamma}(E_{1},T) =12E1n¯F(E1)Tr[(p/1+m1)ImΣ(p1)].\displaystyle=-\frac{1}{2E_{1}}\bar{n}_{F}(E_{1}){\rm Tr}\bigr{[}({p\!\!\!/}_{1}+m_{1}){\rm Im}\Sigma(p_{1})\bigr{]}. (18)

where, p1=(E1,p1)p_{1}=(E_{1},\vec{p}_{1}) and m1m_{1} are the four-momentum and mass of the injected muon, respectively 444The notations for the injected muon are same with the ones for the injected HQ in the elastic process.. The trace term in Eq. 18 was calculated with a resummed photon propagator, which is very similar with the one in QCD 555The structure of the QED HTL-propagator follows Eq. LABEL:eq:PropTL but with opposite signs S. Peigne´\acute{\rm e} and A. Peshier (2008a).. It is realized E. Braaten and M. H. Thoma (1991a, b) that, in the QCD case, the contributions to Γ{\Gamma} can be obtained from the corresponding QED calculations by simply substitution: e2CFg2e^{2}\rightarrow C_{F}g^{2}, where, ee (gg) is the QED (QCD) coupling constant and CF=4/3C_{F}=4/3 is the quark Casimir factor. It yields E. Braaten and M. H. Thoma (1991a)

Tr[(p/1+m1)ImΣ(p1)]\displaystyle{\rm Tr}\bigr{[}({p\!\!\!/}_{1}+m_{1}){\rm Im}\Sigma(p_{1})\bigr{]} (19)
=2CFg2(1+eE1/T)q𝑑ωn¯B(ω)E12E3𝒜,\displaystyle=-2C_{F}g^{2}(1+e^{-E_{1}/T})\int_{q}\int d\omega\;\bar{n}_{B}(\omega)\frac{E_{1}^{2}}{E_{3}}\mathcal{A}\mathcal{B},

with

𝒜(1ω+v1q2E1)ρL(ω,q)+[v1 2(1(v^1q^)2)\displaystyle\mathcal{A}\equiv(1-\frac{\omega+\vec{v}_{1}\cdot\vec{q}}{2E_{1}})\rho_{L}(\omega,q)+\bigr{[}\vec{v}_{1}^{\;2}(1-(\hat{v}_{1}\cdot\hat{q})^{2}) (20)
ωv1qE1]ρT(ω,q)\displaystyle\qquad-\frac{\omega-\vec{v}_{1}\cdot\vec{q}^{\;}}{E_{1}}\bigr{]}\rho_{T}(\omega,q)
n¯F(E3)δ(E1E3ω)nF(E3)δ(E1+E3ω)\displaystyle\mathcal{B}\equiv\bar{n}_{F}(E_{3})\delta(E_{1}-E_{3}-\omega)-n_{F}(E_{3})\delta(E_{1}+E_{3}-\omega) (21)

where, v1=p1/E1\vec{v}_{1}=\vec{p}_{1}/E_{1} denotes the HQ velocity; nB/F(E)=(eE/T1)1n_{B/F}(E)=(e^{E/T}\mp 1)^{-1} indicates the thermal distributions for Bosons/Fermions and n¯B/F1±nB/F\bar{n}_{B/F}\equiv 1\pm n_{B/F} accounts for the Bose-enhancement or Pauli-blocking effect. Here we have used the short notation qd3q(2π)3\int_{q}\equiv\int\frac{d^{3}\vec{q}}{(2\pi)^{3}} for phase space integrals. Taking |p1||\vec{p}_{1}| and m1/3=mQm_{1/3}=m_{Q} are both much greater than the underlying medium temperature, i.e. |p1|,m1/3T|\vec{p}_{1}|,m_{1/3}\gg T, thus, E1/3TE_{1/3}\gg T and nF(E3)n_{F}(E_{3}) is exponentially suppressed and can be dropped. Moreover, the first δ\delta funciton in Eq. LABEL:eq:Trace_Soft3 can be simplified since E3=(p1q)2+m12E1v1qE_{3}=\sqrt{(\vec{p}_{1}-\vec{q})^{2}+m_{1}^{2}}\approx E_{1}-\vec{v}_{1}\cdot\vec{q}. Concerning the second δ\delta funciton, it cannot contribute for ω\omega less than or on the order of TT E. Braaten and M. H. Thoma (1991a), which will be deleted in this work. Finally, Eq. LABEL:eq:Trace_Soft2 and LABEL:eq:Trace_Soft3 can be reduced to

𝒜ρL(ω,q)+v1 2[1(v^1q^)2]ρT(ω,q)\displaystyle\mathcal{A}\equiv\rho_{L}(\omega,q)+\vec{v}_{1}^{\;2}\bigr{[}1-(\hat{v}_{1}\cdot\hat{q})^{2}\bigr{]}\rho_{T}(\omega,q) (22a)
δ(ωv1q).\displaystyle\mathcal{B}\equiv\delta(\omega-\vec{v}_{1}\cdot\vec{q}). (22b)

By substituting Eq. 22a and 22b into Eq. LABEL:eq:Trace_Soft1, one gets

Tr[(p/1+m1)ImΣ(p1)]\displaystyle{\rm Tr}\bigr{[}({p\!\!\!/}_{1}+m_{1}){\rm Im}\Sigma(p_{1})\bigr{]} (23)
=2CFg2E1q𝑑ωn¯B(ω)δ(ωv1q)\displaystyle=-2C_{F}g^{2}E_{1}\int_{q}\int d\omega\;\bar{n}_{B}(\omega)\delta(\omega-\vec{v}_{1}\cdot\vec{q}\;)
{ρL(ω,q)+v1 2[1(v^1q^)2]ρT(ω,q)}.\displaystyle\qquad\biggr{\{}\rho_{L}(\omega,q)+\vec{v}_{1}^{\;2}\bigr{[}1-(\hat{v}_{1}\cdot\hat{q})^{2}\bigr{]}\rho_{T}(\omega,q)\biggr{\}}.

Equation 18 can be rewritten as

Γ(E1,T)\displaystyle{\Gamma}(E_{1},T) (24)
=CFg2q𝑑ωn¯B(ω)δ(ωv1q)\displaystyle=C_{F}g^{2}\int_{q}\int d\omega\;\bar{n}_{B}(\omega)\delta(\omega-\vec{v}_{1}\cdot\vec{q}\;)
{ρL(ω,q)+v1 2[1(v^1q^)2]ρT(ω,q)}\displaystyle\qquad\biggr{\{}\rho_{L}(\omega,q)+\vec{v}_{1}^{\;2}\bigr{[}1-(\hat{v}_{1}\cdot\hat{q})^{2}\bigr{]}\rho_{T}(\omega,q)\biggr{\}}

with the transverse and longitudinal spectral functions are given by the imaginary part of the retarded propagator

ρT/L(ω,q)2ImDT/LR(ω,q).\displaystyle\rho_{T/L}(\omega,q)\equiv 2\cdot ImD_{T/L}^{R}(\omega,q). (25)

We note that, in the weak coupling limit, a consistent method is to use the HTL resummed propagators, which is contributed by the quasiparticle poles and the Landau damping cuts J. P. Blaizot and E. Iancu (2002). In this analysis, we mainly focus on the low frequency excitation (|ω|<q|\omega|<q), where the Landau damping is dominant and the quasiparticle excitation is irrelevant. The resulting spectral function is therefore denoted by ρ\rho in Eq. 25.

The retarded propagator in Eq.25 reads

DT/LR(ω,q)ΔT/L(ω+iη,q),\displaystyle D_{T/L}^{R}(\omega,q)\equiv{\Delta}_{T/L}(\omega+i\eta,q), (26)

which is defined by setting q0=ω+iηq^{0}=\omega+i\eta (η0+)(\eta\rightarrow 0_{+}), i.e. the real energy, for the dressed gluon propagator ΔT/L(q0,q){\Delta}_{T/L}(q^{0},q) J. P. Blaizot and E. Iancu (2002)

ΔT(q0,q)=1(q0)2q2ΠT(x)\displaystyle{\Delta}_{T}(q^{0},q)=\frac{-1}{(q^{0})^{2}-q^{2}-{\Pi}_{T}(x)} (27)
ΔL(q0,q)=1q2+ΠL(x).\displaystyle{\Delta}_{L}(q^{0},q)=\frac{-1}{q^{2}+{\Pi}_{L}(x)}.

The medium effects are embedded in the HTL gluon self-energy

ΠT(x)=mD22[x2+(1x2)Q(x)]\displaystyle{\Pi}_{T}(x)=\frac{m_{D}^{2}}{2}\bigr{[}x^{2}+(1-x^{2})Q(x)\bigr{]} (28)
ΠL(x)=mD2[1Q(x)]\displaystyle{\Pi}_{L}(x)=m_{D}^{2}\bigr{[}1-Q(x)\bigr{]}

where, x=q0/qx=q^{0}/q; Q(x)Q(x) is the Legendre polynomial of second kind

Q(x)=x2lnx+1x1\displaystyle Q(x)=\frac{x}{2}ln\frac{x+1}{x-1} (29)

and mD2m_{D}^{2} is the Debye screening mass squared for gluon

mD2=g2T2(1+Nf6).\displaystyle m_{D}^{2}=g^{2}T^{2}\bigr{(}1+\frac{N_{f}}{6}\bigr{)}. (30)

The coupling constant, gg, is quantified by the two-loop QCD beta-function O. Kaczmarek and F. Zantow (2005)

g2(μ)=2β0ln(μΛQCD)+β1β0ln[2ln(μΛQCD)]\displaystyle g^{-2}(\mu)=2\beta_{0}ln(\frac{\mu}{\Lambda_{QCD}})+\frac{\beta_{1}}{\beta_{0}}ln\bigr{[}2ln(\frac{\mu}{\Lambda_{QCD}})\bigr{]} (31a)
β0=116π2(1123Nf)\displaystyle\qquad\beta_{0}=\frac{1}{16\pi^{2}}(11-\frac{2}{3}N_{f}) (31b)
β1=1(16π2)2(102383Nf)\displaystyle\qquad\beta_{1}=\frac{1}{(16\pi^{2})^{2}}(102-\frac{38}{3}N_{f}) (31c)

where, πTμ3πT\pi T\leq\mu\leq 3\pi T and ΛQCD=261MeV\Lambda_{QCD}=261~{}{\rm MeV}. NfN_{f} is the number of active flavors in the QGP. Finally, for space-like momentum, Eq. 25 can be expressed as

ρT(ω,q)=\displaystyle\rho_{T}(\omega,q)= πωmD22q3(q2ω2){[q2ω2\displaystyle\frac{\pi\omega m_{D}^{2}}{2q^{3}}(q^{2}-\omega^{2})\biggr{\{}\bigr{[}q^{2}-\omega^{2} (32)
+ω2mD22q2(1+q2ω22ωqlnq+ωqω)]2\displaystyle+\frac{\omega^{2}m_{D}^{2}}{2q^{2}}(1+\frac{q^{2}-\omega^{2}}{2\omega q}ln\frac{q+\omega}{q-\omega})\bigr{]}^{2}
+[πωmD24q3(q2ω2)]2}1\displaystyle+\bigr{[}\frac{\pi\omega m_{D}^{2}}{4q^{3}}(q^{2}-\omega^{2})\bigr{]}^{2}\biggr{\}}^{-1}
ρL(ω,q)=\displaystyle\rho_{L}(\omega,q)= πωmD2q{[q2+mD2(1ω2qlnq+ωqω)]2\displaystyle\frac{\pi\omega m_{D}^{2}}{q}\biggr{\{}\bigr{[}q^{2}+m_{D}^{2}(1-\frac{\omega}{2q}ln\frac{q+\omega}{q-\omega})\bigr{]}^{2} (33)
+(πωmD22q)2}1,\displaystyle+\bigr{(}\frac{\pi\omega m_{D}^{2}}{2q}\bigr{)}^{2}\biggr{\}}^{-1},

with which Eq. LABEL:eq:Trace_Soft7 is computable.

The momentum diffusion coefficients can be calculated by subsituting Eqs. LABEL:eq:Trace_Soft7, 32 and 33 back into Eq. 7 and 8, respectively, and then performing the angular integral, yielding555 Note that the spectral functions (Eq. 32 and 33) are odd, and the resulting ρT/L(ω,q)=ρT/L(ω,q)\rho_{T/L}(-\omega,q)=-\rho_{T/L}(\omega,q) are used to obtain Eq. 34 and 35.

κT(E1,T)=\displaystyle\kappa_{T}(E_{1},T)= CFg28π2v10qmax𝑑qq30v1q𝑑ω(1ω2v12q2)\displaystyle\frac{C_{F}g^{2}}{8\pi^{2}v_{1}}\int_{0}^{q_{max}}dq\;q^{3}\int_{0}^{v_{1}q}d\omega\;(1-\frac{\omega^{2}}{v_{1}^{2}q^{2}}) (34)
[ρL(ω,q)+(v12ω2q2)ρT(ω,q)]cothω2T\displaystyle\bigr{[}\rho_{L}(\omega,q)+(v_{1}^{2}-\frac{\omega^{2}}{q^{2}})\rho_{T}(\omega,q)\bigr{]}coth\frac{\omega}{2T}

and

κL(E1,T)=\displaystyle\kappa_{L}(E_{1},T)= CFg24π2v10qmax𝑑qq0v1q𝑑ωω2v12\displaystyle\frac{C_{F}g^{2}}{4\pi^{2}v_{1}}\int_{0}^{q_{max}}dq\;q\int_{0}^{v_{1}q}d\omega\;\frac{\omega^{2}}{v_{1}^{2}} (35)
[ρL(ω,q)+(v12ω2q2)ρT(ω,q)]cothω2T\displaystyle\bigr{[}\rho_{L}(\omega,q)+(v_{1}^{2}-\frac{\omega^{2}}{q^{2}})\rho_{T}(\omega,q)\bigr{]}coth\frac{\omega}{2T}

with v1|v1|v_{1}\equiv|\vec{v}_{1}|. The maximum momentum exchange is qmax=4E1Tq_{max}=\sqrt{4E_{1}T} in the high-energy limit J. D. Bjorken (1982).

Next, we implement the further calculations by performing a simple change of variables,

t=ω2q2<0q2=t1x2\displaystyle t=\omega^{2}-q^{2}<0\quad\quad q^{2}=\frac{-t}{1-x^{2}} (36)
x=ωq<v1ω=xt1x2,\displaystyle x=\frac{\omega}{q}<v_{1}~{}~{}~{}\qquad\quad\omega=x\sqrt{\frac{-t}{1-x^{2}}},

resulting in

dtdx=|(t,x)(q,ω)|dqdω=2(1x2)dqdω.\displaystyle dtdx=\biggr{|}\frac{\partial(t,x)}{\partial(q,\omega)}\biggr{|}dqd\omega=2(1-x^{2})dqd\omega. (37)

Using Eq. LABEL:eq:OmegaQ2TX and 37, one arrives at Eq. 9-12. Similar results can be found in Ref. W. M. Alberico, A. Beraudo, A. De Pace, A. Molinari, M. Monteno, M. Nardi, and F. Prino (2011); A. Beraduo, A. De Pace, W. M. Alberico, and A. Molinari (2009).

Appendix B Derivation of interaction rate and momentum diffusion coefficients in hard collisions

For two-boday scatterings, the transition rate is defined as the rate of collisions with medium parton ii, which changes the momentum of the HQ (parton ii) from p1\vec{p}_{1} (p2\vec{p}_{2}) to p3=p1q\vec{p}_{3}=\vec{p}_{1}-\vec{q} (p4=p2+q\vec{p}_{4}=\vec{p}_{2}+\vec{q}),

ωQi(p1,q,T)=\displaystyle\omega^{Qi}(\vec{p}_{1},\vec{q},T)= p2n(E2)n¯(E3)n¯(E4)vreldσQi(p1,p2p3,p4).\displaystyle\int_{p_{2}}n(E_{2})\bar{n}(E_{3})\bar{n}(E_{4})v_{rel}d\sigma^{Qi}(\vec{p}_{1},\vec{p}_{2}\rightarrow\vec{p}_{3},\vec{p}_{4}). (38)

Typically, one can assume n¯(E3)=1\bar{n}(E_{3})=1 by neglecting the thermal effects on the HQ after scattering. The differential cross section summed over the spin/polarization and color of the final partons and averaged over those of incident partons,

vreldσQi(p1,p2p3,p4)\displaystyle v_{rel}d\sigma^{Qi}(\vec{p}_{1},\vec{p}_{2}\rightarrow\vec{p}_{3},\vec{p}_{4}) (39)
=12E112E2d3p3(2π)32E3d3p4(2π)32E4|2¯|Qi(2π)4\displaystyle=\frac{1}{2E_{1}}\frac{1}{2E_{2}}\frac{d^{3}\vec{p}_{3}}{(2\pi)^{3}2E_{3}}\frac{d^{3}\vec{p}_{4}}{(2\pi)^{3}2E_{4}}\overline{|\mathcal{M}^{2}}|^{Qi}(2\pi)^{4}
δ(4)(p1+p2p3p4)\displaystyle\qquad\delta^{(4)}(p_{1}+p_{2}-p_{3}-p_{4})

where, vrel=((p1p2)2(m1m2)2)/(E1E2)v_{rel}=(\sqrt{(p_{1}\cdot p_{2})^{2}-(m_{1}m_{2})^{2}})/(E_{1}E_{2}) is the relative velocity between the projectile HQ and the target parton. The interaction rate for a given elastic process reads

ΓQi(E1,T)=\displaystyle{\Gamma}^{Qi}(E_{1},T)= d3qωQi(p1,q,T)\displaystyle\int d^{3}\vec{q}~{}\omega^{Qi}(\vec{p}_{1},\vec{q},T) (40)
=(38,LABEL:eq:DiffCroSec)12E1p2n(E2)2E2p312E3p4n¯(E4)2E4\displaystyle\stackrel{{\scriptstyle(\ref{eq:IndiTraRate},\ref{eq:DiffCroSec})}}{{=}}\frac{1}{2E_{1}}\int_{p_{2}}\frac{n(E_{2})}{2E_{2}}\int_{p_{3}}\frac{1}{2E_{3}}\int_{p_{4}}\frac{\bar{n}(E_{4})}{2E_{4}}
|2¯|Qi(2π)4δ(4)(p1+p2p3p4)\displaystyle\quad\overline{|\mathcal{M}^{2}}|^{Qi}(2\pi)^{4}\delta^{(4)}(p_{1}+p_{2}-p_{3}-p_{4})

The momentum diffusion coefficients can be obtained by inserting Eq. 40 into Eq. 7 and 8, yielding

κTQi(E1,T)\displaystyle\kappa^{Qi}_{T}(E_{1},T) (41)
=12E1p2n(E2)2E2p312E3p4n¯(E4)2E4qT 22\displaystyle=\frac{1}{2E_{1}}\int_{p_{2}}\frac{n(E_{2})}{2E_{2}}\int_{p_{3}}\frac{1}{2E_{3}}\int_{p_{4}}\frac{\bar{n}(E_{4})}{2E_{4}}\frac{\vec{q}_{T}^{\;2}}{2}
θ(|t||t|)|2¯|Qi(2π)4δ(4)(p1+p2p3p4)\displaystyle\quad\theta(|t|-|t^{\ast}|)\overline{|\mathcal{M}^{2}}|^{Qi}(2\pi)^{4}\delta^{(4)}(p_{1}+p_{2}-p_{3}-p_{4})
=1256π4|p1|E1|p2|mind|p2|E2n(E2)1cosψ|maxd(cosψ)\displaystyle=\frac{1}{256\pi^{4}|\vec{p}_{1}|E_{1}}\int_{|\vec{p}_{2}|_{min}}^{\infty}d|\vec{p}_{2}|E_{2}n(E_{2})\int_{-1}^{cos\psi|_{max}}d(cos\psi)
tmintdtωminωmaxdωn¯(ω+E2)G(ω)[ω2t(2E1ωt)24p12]|2¯|Qi\displaystyle\quad\int_{t_{min}}^{t^{\ast}}dt\int_{\omega_{min}}^{\omega_{max}}d\omega\frac{\bar{n}({\omega+E_{2}})}{\sqrt{G(\omega)}}\bigr{[}\omega^{2}-t-\frac{(2E_{1}\omega-t)^{2}}{4\vec{p}_{1}^{2}}\bigr{]}\overline{|\mathcal{M}^{2}}|^{Qi}

and

κLQi(E1,T)\displaystyle\kappa^{Qi}_{L}(E_{1},T) (42)
=12E1p2n(p2)2E2p312E3p4n¯(p4)2E4(2E1ωt2|p1|)2\displaystyle=\frac{1}{2E_{1}}\int_{p_{2}}\frac{n(\vec{p}_{2})}{2E_{2}}\int_{p_{3}}\frac{1}{2E_{3}}\int_{p_{4}}\frac{\bar{n}(\vec{p}_{4})}{2E_{4}}\bigr{(}\frac{2E_{1}\omega-t}{2|\vec{p}_{1}|}\bigr{)}^{2}
θ(|t||t|)|2¯|Qi(2π)4δ(4)(p1+p2p3p4)\displaystyle\quad\theta(|t|-|t^{\ast}|)\overline{|\mathcal{M}^{2}}|^{Qi}(2\pi)^{4}\delta^{(4)}(p_{1}+p_{2}-p_{3}-p_{4})
=1512π4|p1|3E1|p2|mind|p2|E2n(E2)1cosψ|maxd(cosψ)\displaystyle=\frac{1}{512\pi^{4}|\vec{p}_{1}|^{3}E_{1}}\int_{|\vec{p}_{2}|_{min}}^{\infty}d|\vec{p}_{2}|E_{2}n(E_{2})\int_{-1}^{cos\psi|_{max}}d(cos\psi)
tmintdtωminωmaxdωn¯(ω+E2)G(ω)(2E1ωt)2|2¯|Qi\displaystyle\quad\int_{t_{min}}^{t^{\ast}}dt\int_{\omega_{min}}^{\omega_{max}}d\omega\frac{\bar{n}({\omega+E_{2}})}{\sqrt{G(\omega)}}(2E_{1}\omega-t)^{2}\overline{|\mathcal{M}^{2}}|^{Qi}

Note that,

  • (1)

    for hard collisions the momentum exchange is constrained by imposing θ(|t||t|)\theta(|t|-|t^{\ast}|) in the first equality of Eq. LABEL:eq:KappaT3 and LABEL:eq:KappaL3;

  • (2)

    the tree level matrix elements squared includes the contributions from the various channels, which are given in Ref. B. L. Combridge (1979):

    • \bullet

      for Q+qQ+qQ+q\rightarrow Q+q (tt-channel only)

      |2¯|Qq(s,t)\displaystyle\overline{|\mathcal{M}^{2}}|^{Qq}(s,t) (43)
      =2Nf2Ncg449(m12u)2+(sm12)2+2m12tt2\displaystyle=2N_{f}\cdot 2N_{c}\cdot g^{4}\frac{4}{9}\frac{(m_{1}^{2}-u)^{2}+(s-m_{1}^{2})^{2}+2m_{1}^{2}t}{t^{2}}
    • \bullet

      for Q+gQ+gQ+g\rightarrow Q+g (tt, ss and uu-channel combined)

      |2¯|Qg(s,t)\displaystyle\overline{|\mathcal{M}^{2}}|^{Qg}(s,t) (44)
      =2(Nc21)g4[2(sm12)(m12u)t2\displaystyle=2(N_{c}^{2}-1)g^{4}\biggr{[}2\frac{(s-m_{1}^{2})(m_{1}^{2}-u)}{t^{2}}
      +49(sm12)(m12u)+2m12(s+m12)(sm12)2\displaystyle\quad+\frac{4}{9}\frac{(s-m_{1}^{2})(m_{1}^{2}-u)+2m_{1}^{2}(s+m_{1}^{2})}{(s-m_{1}^{2})^{2}}
      +49(sm12)(m12u)+2m12(m12+u)(m12u)2\displaystyle\quad+\frac{4}{9}\frac{(s-m_{1}^{2})(m_{1}^{2}-u)+2m_{1}^{2}(m_{1}^{2}+u)}{(m_{1}^{2}-u)^{2}}
      +19m12(4m12t)(sm12)(m12u)+(sm12)(m12u)+m12(su)t(sm12)\displaystyle\quad+\frac{1}{9}\frac{m_{1}^{2}(4m_{1}^{2}-t)}{(s-m_{1}^{2})(m_{1}^{2}-u)}+\frac{(s-m_{1}^{2})(m_{1}^{2}-u)+m_{1}^{2}(s-u)}{t(s-m_{1}^{2})}
      (sm12)(m12u)m12(su)t(m12u)]\displaystyle\quad-\frac{(s-m_{1}^{2})(m_{1}^{2}-u)-m_{1}^{2}(s-u)}{t(m_{1}^{2}-u)}\biggr{]}

    Concerning the degeneracy factors, in Eq. LABEL:eq:MatrixHQq, 2Nf2N_{f} reflects the identical contribution from all light quark and anti-quark flavors, and 2Nc2N_{c} indicates the summing, rather than averaging, over the helicities and colors of the incident light quark, while in Eq. LABEL:eq:MatrixHQg, the factor 2(Nc21)2(N_{c}^{2}-1) denotes the summing over the polarization and colors of the incident gluon. The running coupling constant takes g(μ)2=4παs(μ)g(\mu)^{2}=4\pi\alpha_{s}(\mu), which is given by Eq. 31a with the scale μ=t\mu=\sqrt{-t}.

  • (3)

    with the help of δ\delta-function, we can reduce the integral in Eq. LABEL:eq:KappaT3 and LABEL:eq:KappaL3 from 9-dimension (9D) to 4D in the numerical calculations, by transforming the integration variables from (p2,p3,p4)(\vec{p}_{2},\vec{p}_{3},\vec{p}_{4}) to (|p2|,cosψ,t)(|\vec{p}_{2}|,cos\psi,t), where ψ\psi is the polar angle of p2\vec{p}_{2}; it yields the resutls as shown in the second equality of Eq. LABEL:eq:KappaT3 and LABEL:eq:KappaL3; the relevant limits of integration together with the additional notations are summarized below:

    |p2|min=|t|+(t)2+4m12|t|4(E1+|p1|)\displaystyle|\vec{p}_{2}|_{min}=\frac{|t^{\ast}|+\sqrt{(t^{\ast})^{2}+4m_{1}^{2}|t^{\ast}|}}{4(E_{1}+|\vec{p}_{1}|)} (45)
    cosψ|max=min{1,E1|p1||t|+(t)2+4m12|t|4|p1||p2|}\displaystyle cos\psi|_{max}=min\biggr{\{}1,\frac{E_{1}}{|\vec{p}_{1}|}-\frac{|t^{\ast}|+\sqrt{(t^{\ast})^{2}+4m_{1}^{2}|t^{\ast}|}}{4|\vec{p}_{1}|\cdot|\vec{p}_{2}|}\biggr{\}} (46)
    tmin=(sm12)2s\displaystyle t_{min}=-\frac{(s-m_{1}^{2})^{2}}{s} (47)
    ωmax/min=b±D2a2with\displaystyle\omega_{max/min}=\frac{b\pm\sqrt{D}}{2a^{2}}~{}with (48)
    a=sm12|p1|\displaystyle\quad\quad a=\frac{s-m_{1}^{2}}{|\vec{p}_{1}|} (49)
    b=2tp1 2[E1(sm12)E2(s+m12)]\displaystyle\quad\quad b=-\frac{2t}{\vec{p}_{1}^{\;2}}\bigr{[}E_{1}(s-m_{1}^{2})-E_{2}(s+m_{1}^{2})\bigr{]} (50)
    c=tp1 2{t[(E1+E2)2s]+4p1 2p2 2sin2ψ}\displaystyle\quad\quad c=-\frac{t}{\vec{p}_{1}^{\;2}}\biggr{\{}t\bigr{[}(E_{1}+E_{2})^{2}-s\bigr{]}+4\vec{p}_{1}^{\;2}\vec{p}_{2}^{\;2}sin^{2}\psi\biggr{\}} (51)
    D=b2+4a2c=t[ts+(sm12)2](4E2sinψ|p1|)2\displaystyle\quad\quad D=b^{2}+4a^{2}c=-t\biggr{[}ts+(s-m_{1}^{2})^{2}\biggr{]}\cdot\biggr{(}\frac{4E_{2}sin\psi}{|\vec{p}_{1}|}\biggr{)}^{2} (52)
    G(ω)=a2ω+bω+c\displaystyle G(\omega)=-a^{2}\omega+b\omega+c (53)
  • (4)

    then, one can follow the procedure of Ref. S. Peigne´\acute{\rm e} and A. Peshier (2008a) for the analytical evaluation of ω\omega integral. The obtained results for κT\kappa_{T} and κL\kappa_{L} are shown in Eq. 13 and 14, respectively.

References