-
Heavy quark dipoles are produced in initial parton hard scatterings and then evolve into charmonium eigenstates. The momentum distribution of
$ c\bar c $ dipoles is approximated to be the$ J/\psi $ momentum distribution in proton-proton (pp) collisions. Therefore, in p-Pb collisions, the initial distribution of primordially produced$ c\bar c $ dipoles can be obtained through a superposition of the effective pp collisions [23],$ \begin{aligned}[b] f_{\Psi}({\boldsymbol p},{\boldsymbol x}|{\boldsymbol b}) =& (2\pi)^3\delta(z)T_{p}({\boldsymbol x}_{T})T_{A}({\boldsymbol x}_{T}-{\boldsymbol b}) \\ &\times {\cal R}_{g}(x_g,\mu_{F},{\boldsymbol x}_{T}-{\boldsymbol b}) {{\rm d} \bar \sigma^\Psi_{pp}\over {\rm d}^3{\boldsymbol p}}, \end{aligned} $
(1) where
$\boldsymbol b$ is the impact parameter,${\boldsymbol x}_{T}$ is the transverse coordinate,$T_A({\boldsymbol x}_{T})=\int {\rm d}z\rho_A({\boldsymbol x}_{T},z)$ is the nuclear thickness function, and the nuclear density is taken as the Woods-Saxon distribution.$T_p({\boldsymbol x}_{T})$ is the proton thickness, where proton density is taken as a Gaussian distribution [23]. The width of the Gaussian function is determined with the proton charge radius$ \langle r\rangle_p=0.9 $ fm [41]. The shadowing effect is included with the inhomogeneous modification factor${\cal R}_{g}$ [42] for gluons with the longitudinal momentum$x_g=e^{y}\ E_{T}/\sqrt{s_{NN}}$ and factorization factor$\mu_{F}=E_{T}$ . Transverse energy and momentum rapidity are defined as$E_{T}=\sqrt{m_\Psi^2+{\boldsymbol p}_{T}^2}$ and$y=1/2\ln((E+p_{z})/(E- p_{z}))$ , respectively. The values ofthe gluon shadowing factor$ \mathcal{R}_g $ are obtained using the EPS09model [43]. The effective initial momentum distribution$\dfrac{{\rm d} \bar \sigma^\Psi_{pp}}{ {\rm d}^3{\boldsymbol p}}$ of charmonium in p-Pb collisions includes the Cronin effect [44]. Before two gluons fuse into a heavy quark dipole, they obtain additional transverse momentum via multi-scatterings with the surrounding nucleons. The extra momentum will be inherited by the produced$ c\bar c $ dipole or charmonium states. With the random walk approximation, the Cronin effect is included with the modification in the momentum-differential cross section measured in pp collisions,$ \begin{align} {{\rm d}\bar \sigma_{pp}^{\Psi}\over {\rm d}^3{\boldsymbol p}} = {1\over \pi a_{gN}l}\int {\rm d}^2{\boldsymbol q}_{T} {\rm e}^{-{\boldsymbol q}_{T}^2\over a_{gN}l} {{\rm d}\sigma_{pp}^{\Psi}\over {\rm d}^3{\boldsymbol p}} , \end{align} $
(2) where
$l({\boldsymbol x}_{T})=0.5T_A({\boldsymbol x}_{T})/\rho_A({\boldsymbol x}_{T},z=0)$ is the average path length of a gluon in the nucleus travelling through before scattering with another gluon in the proton to produce a heavy quark dipole at the position${\boldsymbol x}_{T}$ .$ a_{gN} $ represents the extra transverse momentum square in a unit of length of nucleons before the fusion process. Its value is taken to be$ a_{gN}=0.15\ \rm{GeV^2/fm} $ [45]. The charmonium distribution in pp collisions has been measured by the ALICE Collaboration at 2.76 TeV and 7 TeV [46, 47]. With these data, we parametrize the normalized$p_{T}$ distribution of charmonium at$ \sqrt{s_{NN}}=5.02 $ TeV and obtain$ \begin{align} {{\rm d} N_{J/\psi}\over 2\pi p_T {\rm d}{ p_T}} = {(n-1)\over \pi (n-2) \langle p_T^2\rangle_{pp}}\left[1+{p_T^2\over (n-2)\langle p_T^2\rangle_{pp}}\right]^{-n}, \end{align} $
(3) where
$ n=3.2 $ , and the mean transverse momentum square of charmonium is parametrized as$ \langle p_T^2\rangle_{pp}(y)=12.5\times [1- (y/y_{\rm max})^2] \rm{(GeV/c)^2} $ , in which the maximum rapidity of charmonium is defined with$ y_{\rm max}=\ln(\sqrt{s_{NN}}/m_\Psi) $ [48].$ m_\Psi $ is the charmonium mass. -
The heavy quark potential of the
$ c\bar c $ dipole is modified by a hot medium, which affects the evolution of charmonium wave functions [49–51]. Hot medium effects can be included in the Hamiltonian of$ c\bar c $ dipoles. Because a charm quark is heavy compared with the inner movement of charmonium bound states, the relativistic effect is ignored when considering the inner structure of a charmonium. We employ the time-dependent Schrödinger equation to describe the evolution of$ c\bar c $ dipole wave functions with in-medium complex potentials. Assuming the heavy quark-medium interaction is spherical without angular dependence, there is no mixing between charmonium eigenstates with different angular momenta in the wave function of the$ c\bar c $ dipole. The radial part of the$ c\bar c $ dipole wave function in the center of mass frame is separated as follows:$ \begin{align} {\rm i}\,\hbar {\partial \over \partial t}\psi( r, t) = \Bigg[-{\hbar^2\over 2m_\mu}{\partial ^2\over \partial r^2} +V( r, T) + {L(L+1)\hbar^2\over 2 m_\mu r^2}\Bigg]\psi(r,t), \end{align} $
(4) where r is the relative distance between charm and anti-charm quarks, and t is the proper time in the center of mass frame.
$ m_\mu=m_1m_2/(m_1+m_2)=m_c/2 $ is the reduced mass, and$ m_c $ is the charm quark mass.$ \psi(r,t) $ is defined as$ \psi(r,t)=r R(r,t) $ , where$ R(r,t) $ is the radial part of the$ c\bar c $ dipole wave function. The complete wave function of the$ c\bar c $ dipole can be expanded in the eigenstates of the vacuum Cornell potential,$ \Psi(r,\theta, \phi)= \sum_{nlm}c_{nlm}R_{nl}(r)Y_{lm}(\theta, \phi) $ .$ Y_{lm} $ is the spherical harmonics function, and$ L=(0,1,...) $ is the quantum number of the angular momentum. In an ideal fluid with zero viscosity, the heavy quark potential$ V(r,T) $ is radial. There are no transitions between charmonium eigenstates with different angular momenta L. The potential depends on the local temperature of the medium, which is given by the hydrodynamic model in the next section. The radial Schrödinger equation (Eq. (4)) is solved numerically using the Crank–Nicolson method (taking natural units$ \hbar=c=1 $ ). The numerical form of the Schrödinger equation is simplified to$ \begin{align} {\boldsymbol T}_{j,k}^{n+1}\psi_{k}^{n+1} = \mathcal{V}_{j}^{n}. \end{align} $
(5) Here, j and k are the index of rows and columns in the matrix
$\boldsymbol T$ , respectively. The non-zero elements in the matrix are$ \begin{aligned}[b] &{\boldsymbol T}^{n+1}_{j,j}= 2+2a+bV_j^{n+1}, \\ &{\boldsymbol T}^{n+1}_{j,j+1}={\boldsymbol T}^{n+1}_{j+1,j}= -a, \\ &\mathcal{V}_j^n= a\psi_{j-1}^n +(2-2a-bV_j^n)\psi_j^n +a\psi_{j+1}^n , \end{aligned} $
(6) where i is an imaginary number,
$a= {\rm i} \Delta t/(2m_\mu (\Delta r)^2)$ , and$b= {\rm i}\Delta t$ . The subscript j and superscript n in$ \psi_j^n $ represent the coordinate$ r_j=r_0 +j\cdot \Delta r $ and time$ t^n=t_0 +n\cdot\Delta t $ , respectively.$ \Delta r $ and$ \Delta t $ are the steps of the radius and time in the numerical simulation, respectively, and their values are taken to be$ \Delta t=0.001 $ fm/c and$ \Delta r=0.03 $ fm, respectively.$ t_0 $ is the start time of the Schrödinger equation. The matrix${\boldsymbol T}^{n}$ at each time step depends on the in-medium heavy quark potential$ V(r,T) $ , which is given later.The Schrödinger equation (Eq. (4)) describes the evolution of the wave function of the
$ c\bar c $ dipole from$ t\ge t_0 $ . The initial wave function of the$ c\bar c $ dipole is taken to be one of the charmonium eigenstates. After traveling through the hot medium, the fractions$ |c_{nl}(t)|^2 $ of each charmonium eigenstate (1S, 1P, 2S, etc.) in the$ c\bar c $ dipoles change with time.$ c_{nl}(t) $ is defined as$ \begin{align} c_{nl}(t) &= \int R_{nl}(r) {\rm e}^{-{\rm i}E_{nl} t} \psi(r,t) r {\rm d} r , \end{align} $
(7) where the radial wave function
$ \psi(r,t) $ is given by Eq. (5). The ratio of the final and initial fractions of a certain charmonium state in one$ c\bar c $ dipole is expressed as$ R^{\rm direct}(t) ={|c_{nl}(t)|^2\over |c_{nl}(t_0)|^2} $ . In p-Pb collisions, the initial spatial and momentum distributions of primordially produced$ c\bar c $ dipoles are given by Eq. (1). After averaging over the position and momentum bins of different$ c\bar c $ dipoles in p-Pb collisions, we can obtain the ensemble-averaged fractions of a certain charmonium state in the$ c\bar c $ dipole$ \langle |c_{nl}(t)|^2\rangle_{\rm en} $ . The direct nuclear modification factor of the charmonium eigenstate ($ n,l $ ) is written as$ \begin{aligned}[b] R_{pA}^{\rm direct}(nl) &={\langle |c_{nl}(t)|^2\rangle_{\rm en}\over \langle |c_{nl}(t_0)|^2\rangle_{\rm en}} \\ &={\displaystyle\int {\rm d}{\boldsymbol x}_{\Psi}{\rm d}{\boldsymbol p}_{\Psi} |c_{nl}(t, {\boldsymbol x}_{\Psi}, {\boldsymbol p}_{\Psi})|^2{{{\rm d}N^{\Psi}_{pA}}\over{\rm d}{\boldsymbol x}_{\Psi} {\rm d}{\boldsymbol p}_{\Psi}} \over \displaystyle\int {\rm d}{\boldsymbol x}_{\Psi} {\rm d} {\boldsymbol p}_{\Psi} |c_{nl}(t_0,{\boldsymbol x}_0, {\boldsymbol p}_{\Psi})|^2 {\overline {{\rm d}N^{\Psi}_{pA}}\over {\rm d}{\boldsymbol x}_{\Psi}{\rm d}{\boldsymbol p}_{\Psi}}}, \end{aligned} $
(8) where
${\boldsymbol x}_{\Psi}$ and${\boldsymbol p}_{\Psi}$ are the position and total momentum of the correlated$ c\bar c $ dipole, respectively. Without the hot medium effects, these correlated$ c\bar c $ dipoles are simply charmonium eigenstates without dissociation.${{\rm d}N_{pA}^{\Psi}\over {\rm d}{\boldsymbol x}_{\Psi}{\rm d}{\boldsymbol p}_{\Psi}}$ is the initial spatial and momentum distributions of primordially produced charmonium in p-Pb collisions and is given by Eq. (1). Note that in the denominator,${\overline{{\rm d}N_{pA}^{\Psi}}\over {\rm d}{\boldsymbol x}_{\Psi}{\rm d}{\boldsymbol p}_{\Psi}}$ is calculated using Eq. (1) excluding cold nuclear matter effects.After considering feed-down contributions from excited states, the nuclear modification factor of
$ J/\psi $ can be obtained (which is given in the experimental data),$ \begin{align} R_{pA}(J/\psi) = {\sum_{nl} \langle |c_{nl}(t)|^2\rangle_{\rm en} f_{pp}^{nl} \mathcal{B}_{nl\rightarrow J/\psi}\over \sum_{nl} \langle |c_{nl}(t_0)|\rangle^2\rangle_{\rm en} f_{pp}^{nl} \mathcal{B}_{nl\rightarrow J/\psi}} ,\end{align} $
(9) where
$ \mathcal{B}_{nl\rightarrow J/\psi} $ is the branching ratio of charmonium eigenstates with the quantum number$ (n,l) $ decaying into the ground state$ J/\psi $ . We consider the decay channels$ \chi_c\rightarrow J/\psi $ and$ \psi(2S)\rightarrow J/\psi $ .$ f_{pp}^{nl} $ is the direct production of the charmonium eigenstate ($ J/\psi $ ,$ \chi_c $ ,$ \psi(2S) $ ) without the feed-down process in pp collisions. The ratio of direct charmonium production is extracted to be$ f_{pp}^{J/\psi}:f_{pp}^{\chi_c}:f_{pp}^{\psi(2S)} =0.68:1:0.19 $ [52]. -
In vacuum, the heavy quark potential in the quarkonium can be approximated as the Cornell potential. At finite temperature, the Cornell potential is screened by thermal light partons. The real part of the in-medium heavy quark potential is between the limits of the free energy F and internal energy U of charmonium. The in-medium potential has been studied using lattice QCD calculations and potential models [53–56]. We parametrize the temperature and coordinate dependence of free energy using the formula
$ \begin{aligned}[b] F(T,r) =& -{\alpha\over r}[e^{-\mu r}+\mu r] \\ & -{\sigma \over 2^{3/4}\Gamma[3/4]}\left({r\over \mu}\right)^{1/2} K_{1/4}[(\mu r)^2] +{\sigma\over 2^{3/2}\mu }{\Gamma[1/4]\over \Gamma[3/4]}, \end{aligned} $
(10) where
$ \alpha=\pi/12 $ and$ \sigma=0.2\ \rm{GeV^2} $ are given in the Cornell potential$ V_c(r)={-\alpha/r}+\sigma r $ . The Γ and$ K_{1/4} $ are the Gamma function and modified Bessel function, respectively. The screened mass in Eq. (10) is taken as [54]$ \begin{aligned}[b] {\mu(\bar T)\over \sqrt{\sigma}} = s\bar{T} +a \sigma_t \sqrt{\pi \over 2} \left[\mathrm{erf}\left({b\over \sqrt{2}\sigma_t}\right) - \mathrm{erf}\left({b-\bar{T}\over \sqrt{2}\sigma_t}\right)\right], \end{aligned} $
(11) with
$ {\bar T}\equiv T/T_c $ , where$ T_c $ is the critical temperature of the deconfined phase transition. Other parameters are taken as$ s=0.587 $ ,$ a=2.150 $ ,$ b=1.054 $ , and$ \sigma_t=0.07379 $ .$ \mathrm{erf}(z) $ is the error function. The internal energy of a heavy quarkonium can be obtained via the relation$ U(T,r)= F+ T(-\partial F/\partial T) $ . When the slope of the line becomes flat, this indicates that there is no attractive force to restrain the wave function at the distance r. At temperatures of approximately$ T_c $ , there is a sudden shift in the screened mass$ \mu(\bar T) $ [54]. The internal energy may become slightly larger than the vacuum Cornell potential. This behavior can be seen in$ U(T,r) $ at$ r\sim 0.4 $ fm in Fig. 1 and becomes more evident at$ T\rightarrow T_c $ . To avoid this subtlety, for the heavy quark potential, we take the free energy as the limit of strong color screening and the vacuum Cornell potential as the limit of extremely weak color screening. The realistic potential is between these two limits. Different heavy quark potentials in Fig. 1 are inserted into the Schrödinger equation to calculate the nuclear modification factors of$ J/\psi $ and$ \psi(2S) $ in the next section.Figure 1. (color online) Different parametrizations of the real part of heavy quark potentials as a function of r at
$ T=1.5T_c $ . The free energy$ F(r,T) $ , internal energy$ U(r,T) $ , and Cornell potential$ V_c(r) $ are plotted with different colored lines.In the hot medium, quarkonium bound states can also be dissociated by inelastic scatterings with thermal light partons. This process contributes an imaginary part to the potential
$ V(T,r) $ . We parameterize the temperature and spatial dependence of the imaginary potential using$ \begin{align} &V_I(T,\bar r)= -{\rm i}\,T(a_1\, {\bar r} + a_2 {\bar r}^2)\,, \end{align} $
(12) where i is the imaginary unit, and
$ \bar r\equiv r/{\rm fm} $ is a dimensionless variable. The dimensionless coefficients$ a_1 $ and$ a_2 $ are obtained by invoking Bayesian inference to fit the lattice QCD calculations [57]. We focus on the temperature relevant to p-Pb collisions,$ T_c <T< 1.9\; T_c $ . The results are shown in Fig. 2, where the gray band represents the$ 95\% $ confidence interval, and the black curve corresponds to the parameter set$ a_1=-0.040 $ and$ a_2=0.50 $ , which maximizes the posterior distribution. In$ V_I $ , the magnitude of the imaginary potential becomes smaller at smaller distances. This results in a weaker reduction in the$ J/\psi $ component than the$ \psi(2S) $ component in the wave function of the$ c\bar c $ dipole. Because the imaginary potential in Fig. 2 is calculated in the gluonic medium, we take the same formula for QGP in heavy-ion collisions, which contributes some uncertainty to the suppression of charmonium in p-Pb collisions [56, 58]. The uncertainty on the imaginary potential is partially considered with the theoretical band in Fig. 2, which will be reflected in the charmonium$ R_{pA} $ .Figure 2. (color online) Imaginary part of the heavy quark potential as a function of distance. The gray band represents the
$95$ % confidence region, whereas the black curve corresponds to the maximum a posteriori parameter set. The data is cited from [57]. Symbols from purple to red correspond to results from low to high temperature.In the hot medium produced in p-Pb collisions, heavy quark dipoles experience different local temperatures as they move along different trajectories. The real and imaginary parts of the potential, depending on the local temperatures, also change with time. The wave package at each time step is obtained from the Schrödinger equation, while its normalization is reduced by the imaginary part of the Hamiltonian. Therefore, the fractions of charmonium eigenstates in the wave package change with time owing to in-medium potentials.
-
The dynamical evolution of the hot medium produced in p-Pb collisions at
$ \sqrt{s_{NN}}=5.02 $ TeV is described by hydrodynamic equations [21].$ \begin{align} \partial_{\mu\nu} T^{\mu\nu}=0 ,\end{align} $
(13) where
$ T^{\mu\nu}=(e+p)u^\mu u^\nu-g^{\mu\nu}p $ is the energy-momentum tensor, e and p are the energy density and pressure, respectively, and$ u^\mu $ is the four velocity of the medium. The equation of state is required to close the hydrodynamic equations. The deconfined phase is treated as an ideal gas of gluons and massless u and d quarks plus s quarks with the mass$ m_s=150 $ MeV. The confined phase is treated using the hadron resonance gas model (HRG) [59]. Two phases are connected with a first-order phase transition, and the critical temperature of the phase transition is determined as$ T_c=165 $ MeV by choosing the mean field repulsion parameter and bag constant to be$ K=450 \ \rm{MeV\,fm^3} $ and$ B^{1/4}=236 $ MeV [60], respectively. With the multiplicity of light hadrons measured in p-Pb collisions and theoretical simulations from other hydrodynamic models [21, 61], we take the maximum initial temperature of the hot medium to be$T_0({\boldsymbol x}_T=0|b=0)= 248$ MeV in forward rapidity and$ 289 $ MeV in backward rapidity. Event-by-event fluctuations in hydrodynamic evolution are not yet included. The profile of the initial energy density is also consistent with the results from a multiple phase transport (AMPT) model [62].Hydrodynamic equations begin evolution from
$ \tau_0=0.6 $ fm/c, where the hot medium is assumed to reach local equilibrium. The time evolution of the local temperature at$ { x}_T=0 $ in forward and backward rapidity at most central collisions with the impact parameter b =0 is plotted in Fig. 3. Medium evolution with other impact parameters can be obtained via the scale of initial entropy, which depends on$ N_p(b) $ and$N_{\rm coll}(b)$ .
Investigating color screening in proton-nucleus collisions with complex potentials
- Received Date: 2022-05-23
- Available Online: 2022-11-15
Abstract: Color screening and parton inelastic scattering modify the heavy-quark antiquark potential in mediums consisting of particles from quantum chromodynamics (QCD), leading to the suppression of quarkonium production in relativistic heavy-ion collisions. Owing to the small charm/anti-charm (