Further developments of a multi-phase transport model for relativistic nuclear collisions

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Further developments of a multi-phase transport model for relativistic nuclear collisions

Zi-Wei Lin
Liang Zheng
Nuclear Science and TechniquesVol.32, No.10Article number 113Published in print 01 Oct 2021Available online 11 Oct 2021
8900

A multi-phase transport (AMPT) model was constructed as a self-contained kinetic theory-based description of relativistic nuclear collisions as it contains four main components: the fluctuating initial condition, a parton cascade, hadronization, and a hadron cascade. Here we review the main developments after the first public release of the AMPT source code in 2004 and the corresponding publication that described the physics details of the model at that time. We also discuss possible directions for future developments of the AMPT model to better study the properties of the dense matter created in relativistic collisions of small or large systems.

QGPTransport modelHeavy-ion collisions

1 Introduction

In high energy heavy ion collisions [1], a hot and dense matter made of parton degrees of freedom, the quark-gluon plasma (QGP), has been expected to be created [2]. Experimental data from the Relativistic Heavy Ion Collider (RHIC) and the Large Hadron Collider (LHC) [3-8] strongly indicate that the QGP is indeed created in heavy ion collisions at high energies [9]. Comprehensive comparisons beween the experimental data and theoretical models are essential for the extraction of key properties of the high density matter, including the structure of the QCD phase diagram at high temperature and/or high net-baryon density. Many theoretical models including transport models [10-14], hydrodynamic models [15-18], and hybrid models [19-21] have been constructed to simulate and study the phase space evolution of the QGP.

A multi-phase transport (AMPT) model [13] is one such model. The AMPT model aims to apply the kinetic theory approach to describe the evolution of heavy-ion collisions as it contains four main components: the fluctuating initial condition, partonic interactions, hadronization, and hadronic interactions. The default version of the AMPT model [11, 22] was first constructed. Its initial condition is based on the Heavy Ion Jet INteraction Generator (HIJING) two-component model [23, 24], then minijet partons enter the parton cascade and eventually recombine with their parent strings to hadronize via the Lund string fragmentation [25]. The default AMPT model can well describe the rapidity distributions and transverse momentum (pT) spectra of identified particles observed in heavy ion collisions at SPS and RHIC. However, it significantly underestimates the elliptic flow (v2) at RHIC.

Since the matter created in the early stage of high energy heavy ion collisions is expected to have a very high energy density and thus should be in parton degrees of freedom, the string melting version of the AMPT (AMPT-SM) model [26] was then constructed, where all the excited strings from a heavy ion collision are converted into partons and a spatial quark coalescence model is invented to describe the hadronization process. String melting increases the parton density and produces an over-populated partonic matter [27], while quark coalescence further enhances the elliptic flow of hadrons [26, 28]. As a result, the string melting AMPT model is able to describe the large elliptic flow in Au+Au collisions at RHIC energies with a rather small parton cross section [26, 29].

The source code of the AMPT model was first publicly released online around April 2004, and a subsequent publication [13] provided detailed descriptions of the model such as the included physics processes and modeling assumptions. The AMPT model has since been widely used to simulate the evolution of the dense matter created in high energy nuclear collisions. In particular, the string melting version of the AMPT model [13, 26] can well describe the anisotropic flows and particle correlations in collisions of small or large systems at both RHIC and LHC energies [13, 26, 30-33]. The AMPT model is also a useful test bed of different ideas. For example, the connection between the triangular flow and initial geometrical fluctuations was discovered with the help of AMPT simulations [34], and the model has also been applied to studies of vorticity and polarization in heavy ion collisions [35-37].

Experimental data from heavy ion collisions fit with hydrodynamics-inspired models suggest that particles are locally thermalized and possess a common radial flow velocity [38]. Large momentum anisotropies such as the elliptic flow [39] have been measured in large collision systems, as large as the hydrodynamics predictions [7, 40]. This suggests that the collision system is strongly interacting and close to local thermal equilibrium [9]. Transport models can also generate large anisotropic flows. The string melting AMPT model [13, 26] can describe the large anisotropic flows with a rather small parton cross section of ˜3 mb [26] and the flow enhancement from quark coalescence [26, 28, 29, 41, 42]. Without the quark coalescence, a pure parton transport for minijet gluons requires an unusually large parton cross section of ˜40- -50 mb [29, 43] for the freezeout gluons to have a similar magnitude of elliptic flow as charged hadrons in the experiments. This minijet gluon system, despite a factor of ˜2.5 lower parton multiplicity at mid-rapidity, has a factor of ˜6 smaller mean free path than the string melting AMPT model for 200A GeV Au+Au collisions at impact parameter b=8 fm [29]. In general, for large systems at high energies transport models tend to approach hydrodynamics since the average number of collisions per particle is large and thus the bulk matter is close to local equilibrium. Hydrodynamics models and transport models are also complementary to each other. For example, hydrodynamics models provide a direct access to the equation of state and transport coefficients, while transport models can address non-equilibrium dynamics and provide a microscopic picture of the interactions.

Recent data from small systems, however, hint at significant anisotropic flows in high multiplicity pp and pPb collisions at the LHC [44] and p/d/3He+Au collisions at RHIC [45, 46]. Hydrodynamic calculations seem to describe the experimental data well [47, 48]. The AMPT-SM model also seems to describe the measured correlations [30]. This suggests that the collision of these small systems might create a QGP as well, in contrast to naive expectations. On the other hand, it is natural to expect hydrodynamic models and transport models to be different for small colliding systems due to non-equilibrium dynamics. Indeed, recently it has been realized that parton transport can convert the initial spatial anisotropies into significant anisotropic flows in the momentum space through the parton escape mechanism [49, 50], especially in small systems where the average number of collisions per particle is small. Kinetic theory studies also show that a single scattering is very efficient in changing the particle momentum distribution [51]. There are also many studies on whether and how hydrodynamics could be applicable to small systems [52, 53]. In addition, there are active debates on whether the momentum anisotropies in small systems mainly come from initial state correlations [54, 55] or final state interactions [49-51, 56, 57]. Furthermore, the differences between the anisotropic flow data of small systems from different collaborations still need to be fully resolved [46, 58, 59]. Therefore, the system size dependence of various observables, particularly the anisotropic flows from small to large systems, could provide key information on the origin of collectivity.

The paper is organized as follows. After the introduction, in Sec. 2 we review the main developments of the AMPT model after the first public release of its source code in 2004 [13, 60, 61]. They include the addition of deuteron productions in Sec. 2.1, the string melting model that can simultaneously reproduce the yield, transverse momentum spectra and elliptic flow of the bulk matter in heavy ion collisions in Sec. 2.2, the new quark coalescence model in Sec. 2.3, incorporation of the finite nuclear thickness along beam directions in Sec. 2.4, incorporation of modern parton distribution functions of nuclei in Sec. 2.5, improved treatment of heavy quark productions in Sec. 2.6, the introduction of local nuclear scaling of key input parameters to describe the system size dependence in Sec. 2.7, incorporation of PYTHIA8 and nucleon substructure in the initial condition in Sec. 2.8, and benchmark and improvement of the parton cascade algorithm in Sec. 2.9. We then briefly review other developments of the AMPT model in Sec. 3. Finally, in Sec. 4 we summarize and discuss possible directions for further developments of the AMPT model.

2 Main developments

We now review the main developments of the AMPT model after the first public release of the AMPT source code in 2004 [60, 61] and the corresponding publication that described the physics details of the model at that time [13]. These developments are listed mostly in chronological order. In terms of the four main components of the AMPT model, Sections 2.2, 2.4, 2.5, 2.6, 2.7, 2.8 are about the initial condition, Sec. 2.9 is about the parton cascade, Sec. 2.3 is about the hadronization, while Sec. 2.1 is about the hadron cascade. Currently the public versions of the AMPT model since v1.26t5/v2.26t5 [61] have incorporated the changes made in the developments described in Sections 2.1 and 2.2; changes from the other developments will be released in the future.

2.1  Deuteron productions in the hadron cascade

Light nuclei such as deuteron (d) and triton (t) are produced and observed in high energy nuclear collisions at RHIC and LHC [62, 63]. They have been proposed to be important for the search of the QCD critical point [64-67] and thus the study of light nuclei has become more active recently. Currently the production mechanism of light nuclei is still under debate, as there are several different models including the statistical model [68, 69], the nucleon coalescence model [70-74], and dynamical models based on the kinetic theory [75-77].

We have modified the AMPT model to provide a kinetic theory description of deuteron production and annihilation by adding the following reactions [77]:

BBMd, (1)

where M = π, ρ, ω, and η, while B and B’ stand for baryons N, Δ, P11(1440), and S11(1535). Note that the hadron cascade component of the AMPT model [13], based on a relativistic transport (ART) model [84-86], already includes the interactions of π, K, η, ρ, ω, ϕ, K*, N, Δ(1232), P11(1440), S11(1535) as well as their antiparticles. For the cross sections of the reactions BBMd, we assume that their angular integrated mean squared matrix elements that are averaged over initial and summed over final spins and isospins are the same as that for the reaction NNdπ at same center of mass energy s. The cross sections for the inverse reactions MdBB are then determined from the detailed balance. In addition to the production and annihilation processes for deuterons, we also include their elastic scatterings with mesons M and baryons B [77].

Experimentally, the cross sections for both the reaction ppdπ+ [78-80] and the reaction π+dpp [81-83, 87] have been extensively measured, and the former is given by

σ(ppdπ+)=14pπpNf(s), (2)

where pN and pπ are, respectively, the magnitude of the three-momenta of initial and final particles in the center of mass frame. The function f(s), which is proportional to the angular integrated mean squared matrix elements that are summed over initial and final spins for the reaction ppπ+d, is parameterized as

pic (3)

where s is in the unit of GeV and f(s) is in the unit of mb. For the inverse reaction dπ+, its cross section is related to that for ppdπ+ via the detailed balance relation:

σ(dπ+pp)=2pN23pπ2σ(ppdπ+). (4)

These parameterizations are compared with the experimental data in Fig. 1. The cross sections for the isospin averaged reactions NNdπ and πd can then be obtained as σ(NNdπ)=3σ(pp+)/4 and σ(dπNN)=σ(dπ+pp).

Fig. 1
(Color online) Experimental data on the total cross sections of ppdπ+ [78-80] (filled symbols) and dπ+pp [81-83] (open symbols) in comparison with our parameterizations (solid curves).
pic

We have coupled the above deuteron transport with an initial hadron distribution after hadronization as parameterized by a blast wave model [77], where a nucleon coalescence model using the deuteron Wigner function [88] was also applied for comparison. We find that the transport model gives very similar deuteron pT spectra as the coalescence model; however the elliptic flows from the two models are different. In particular, the transport model gives a deviation of the elliptic flow from the exact nucleon number scaling at relatively high pT and agrees better with the measured data.

On the other hand, the deuteron yield obtained directly from the AMPT-SM model is typically much lower than the experimental data. This could be due to the assumed relation between the BBMd and ppdπ cross sections, which can be further constrained by using the measured total π d cross section, or the lack of additional production channels such as πnpπd [89]. The low yield could also be partly due to the assumption of no primordial deuteron formation from quark coalescence. There are also studies [73, 90] that applied the nucleon coalescence model to the kinetic freezeout nucleon distributions from the AMPT-SM model. It has been found that the resultant light nuclei yields depend sensitively on the freezeout surface, which is affected by both the partonic expansion and the hadronization (quark coalescence) criterion. The yields also depend on the coalescence function used for the light nuclei [90], especially for small collision systems where the suppression due to the light nuclei size [91] could be significant. Further improvements of the AMPT model regarding the deuteron cross sections, the parton phase, and the hadronization criterion will benefit the studies of light nuclei.

2.2  String melting model to describe the bulk matter

The Lund string model [25] is used in both the default and string melting versions of the AMPT model. In the default AMPT model, minijet partons recombine with their parent strings after the parton cascade to hadronize via the Lund string model into primordial hadrons. In the AMPT-SM model, the primordial hadrons that would be produced from the excited Lund strings in the HIJING model are “melt” into primordial quarks and antiquarks. Therefore, the parameters in the Lund string model affect the AMPT model results. In the Lund model, one assumes that a string fragments into quark-antiquark pairs with a Gaussian distribution in transverse momentum. Hadrons are formed from these quarks and antiquarks, with the longitudinal momentum given by the Lund symmetric fragmentation function [92, 93]

f(z)z1(1z)aL ebLmT2/z. (5)

In the above, z represents the light-cone momentum fraction of the produced hadron with respect to that of the fragmenting string and mT is the transverse mass of the hadron.

When using the HIJING values [23, 24] for the key Lund string fragmentation parameters, aL=0.5 and bL=0.9 GeV-2, the default AMPT model works well for particle yields and pT spectra in pp collisions at various energies. However, it gives too small a charged particle yield in central Pb+Pb collisions at the SPS energy of ELAB=158A GeV [11, 22]. Instead, modified values of aL=2.2 and bL=0.5 GeV-2 were needed to fit the charged particle yield and pT spectra in Pb+Pb collisions at SPS. For heavy ion collisions at higher energies such as RHIC energies, the default version of the AMPT model with these parameter values also reasonably describes hadron dN/dη, dN/dy and the pT spectra in heavy ion collisions, although it underestimates the elliptic flow [26].

On the other hand, the AMPT-SM model [26, 13], due to its dense parton phase and quark coalescence, reasonably describes the elliptic flow [26] and two-pion interferometry [94] in heavy ion collisions. However, the versions before 2015 [61] (i.e., before v2.26t5) could not reproduce well the hadron dN/dη, dN/dy and pT spectra (when using the same Lund parameters as the default version). For example, they overestimated the charged particle yield and significantly underestimated the slopes of the pT spectra [13]. In an earlier attempt to reproduce data in Pb+Pb collisions at LHC energies with the AMPT-SM model, the default HIJING values for the Lund string fragmentation parameters were used [95] together with the strong coupling constant αs=0.33 (instead of 0.47); there the model reasonably reproduced the yield and elliptic flow of charged particles but underestimated the pT spectrum (except at low pT).

It was later realized that this problem of the AMPT-SM model can be solved [27] by using a very small value for the Lund fragmentation parameter bL together with an upper limit on strange quark productions. Then the AMPT-SM model can reasonably reproduce the pion and kaon yields, pT spectra, and elliptic flows at low pT (below ˜1.5 GeV/c) in central and semi-central Au+Au collisions at the RHIC energy of sNN=200 GeV and Pb+Pb collisions at the LHC energy of 2.76 TeV [27]. In particular, we found that bL=0.15 GeV-2 is needed [27], which is much lower than in the value used in previous studies [11, 13, 22, 26, 95]. Note that, for a smaller bL value, the effective string tension κ, as given by [22, 13]

κ1bL(2+aL), (6)

is higher and thus gives a larger mean transverse momentum for the initial quarks after string melting. In addition, the AMPT model assumes that the relative production of strange to non-strange quarks increases with the effective string tension [22, 13]. This is because the quark-antiquark pair production from string fragmentation in the Lund model is based on the Schwinger mechanism [96], where the production probability is proportional to exp(πm2/κ) at transverse mass m. As a result, the strange quark suppression relative to light quarks, exp[π(ms2mu2)/κ], is reduced for a higher string tension. It is found that an upper limit of 0.40 on the relative production of strange to non-strange quarks is needed for the AMPT-SM model [27].

Figure 2 shows the AMPT-SM results of pions and kaons for central (b<3 fm) and mid-central (b=7.3 fm) [97] Au+Au events at 200A GeV as well as central (b<3.5 fm) and mid-central (b=7.8 fm) [98] Pb+Pb events at 2.76A TeV. Also plotted for comparisons are the corresponding data for 0-5% and 20-30% centralities on dN/dy [99-101] in panels (a) and (d), the pT spectra at mid-rapidity for the 0-5% centrality in panels (c) and (e), and v2{EP} at mid-rapidity for the 20-30% centrality in panels (d) and (f). We see good agreements between the model results and the dN/dy data in both central and mid-central events at RHIC and LHC energies. The value of 0.55 is used for aL at the top RHIC energy, while the value of 0.30 is used at the LHC energy since it gives a slightly better fit of the ALICE data [101] than the value of 0.55. We also see that the model roughly reproduces the observed pT spectra at mid-rapidity below ˜2 GeV/c. In addition, the AMPT-SM model roughly describes the pion and kaon elliptic flow data on v2{EP} [102, 103] at low pT.

Fig. 2
(Color online) AMPT-SM results for pions (upper panels) and kaons (lower panels) on dN/dy of (a) π+ and (d) K+ in central and mid-central collisions, pT spectra dN/(2πpTdpTdy) in the unit of c2/GeV2 of (b) π+ and (e) K+ at mid-rapidity in central collisions, and elliptic flow v2{EP} of (c) charged pions and (f) charged kaons at mid-rapidity in mid-central collisions in comparison with the experimental data for central (0-5%) and/or mid-central (20-30%) Au+Au collisions at 200A GeV and Pb+Pb collisions at 2.76A TeV.
pic

This choice of settings for the AMPT-SM model [27] reasonably and simultaneously reproduces the particle yield, pT spectra and elliptic flow of the bulk matter in central and semi-central AA collisions. Therefore, it enables us to make more reliable studies, such as the calculation of the evolution of energy density, effective temperatures, and transverse flow of the parton phase [27], and comprehensive predictions for Pb+Pb collisions at the top LHC energy of 5.02 TeV [31].

An example of the 5.02 TeV predictions from the AMPT-SM version v2.26t5 [31] is shown in Fig. 3, where the results on the η dependence of elliptic flow are shown in panels (a) and (b) for two centralities and the results on the factorization ratio r2(ηa,ηb) are shown in panels (c) to (f) for four centralities. We see that the AMPT-SM model reasonably reproduces the observed v2(η) magnitudes and shapes at 15-25% and 25-50% centralities from CMS [104] (filled circles) and ATLAS [105] (open circles) for Pb+Pb collisions at 2.76 TeV and from PHOBOS [106] (open diamonds) for Au+Au collisions at 200 GeV. We also see that the AMPT results on the factorization ratio r2(ηa,ηb) as a function of ηa at 2.76 TeV are rather consistent with the corresponding CMS data [107], similar to a study [108] that used the AMPT-SM model as the initial condition for an ideal (3+1)D hydrodynamics. Furthermore, the AMPT-SM results show that the longitudinal correlation is much suppressed in Au+Au collisions at 200 GeV but slightly enhanced in Pb+Pb collisions at 5.02 TeV. Note that the longitudinal correlation comes naturally in the AMPT-SM model since each excited string typically produces many initial partons over a finite η range. Therefore, the initial transverse spatial geometry of the parton matter including the event plane has a strong correlation over a finite η range, and through partonic and hadronic interactions, the azimuthal anisotropies vn will then develop longitudinal correlations.

Fig. 3
(Color online) AMPT-SM results on the η dependence of v2 in comparison with data for (a) the 15-25% centrality and (b) the 25-50% centrality, and (c)-(f) AMPT-SM results on the factorization ratio r2(ηa,ηb) as functions of ηa in comparison with the CMS data for different centralities.
pic

We note that the AMPT model may not be reliable at higher pT, as indicated by Fig. 2, since it lacks inelastic parton collisions [13, 109] and consequently the radiative energy loss of partons that is important for high pT partons. In addition, the string melting AMPT model up to now uses quark coalescence to model the hadronization of all partons, while the hadronization of high pT partons and partons far away from their coalescing partners should be treated differently, e.g., with independent fragmentation [110] or string fragmentation [111].

2.3  Improved quark coalescence

After parton scatterings, a spatial quark coalescence model is used to describe the hadronization process in the AMPT-SM model. It combines a quark with a nearby antiquark to form a meson and combines three nearby quarks (or antiquarks) into a baryon (or an antibaryon). For quarks and antiquarks in an event, the original quark coalescence model in AMPT [13, 26, 27, 31] searches for a meson partner before searching for baryon or antibaryon partners. Specifically, each quark (or antiquark) has its default coalescence partner(s), which are just the one or two valence parton(s) from the decomposition of the quark’s parent hadron from the string melting process. Then for any available (i.e., not-yet-coalesced) quark (or antiquark) that originally comes from the decomposition of a meson, the quark coalescence model searches all available antiquarks (or quarks) and selects the closest one in distance (in the rest frame of the quark-antiquark system) as the new coalescence partner to form a meson. After these meson coalescences are all finished, for each remaining quark (or antiquark) the model searches all available quarks (or antiquarks) and selects the closest two in distance as the new coalescence partners to form a baryon (or an antibaryon). As a result, the total number of baryons in any event after quark coalescence is the same as the total number before. Similarly, the quark coalescence process also conserves the number of antibaryons and the number of mesons in an event.

However, this separate conservation of the numbers of baryons, antibaryons, and mesons through the quark coalescence for each event is unnecessary, because only conserved charges such as the number of net-baryons and the number of net-strangeness need to be conserved. Therefore, we improved the coalescence method [32, 112] by removing the constraint that forced the separate conservations. Specifically, for any available quark, the new coalescence model searches all available antiquarks and records the closest one in relative distance (denoted as dM) as the potential coalescence partner to form a meson. The model also searches all available quarks and records the closest one in distance as a potential coalescence partner to form a baryon, and then searches all other available quarks again and records the one that gives the smallest average distance (i.e. the average of the three relative distances among these three quarks in the rest frame of the three-quark system, denoted as dB) as the other potential coalescence partner to form a baryon.

In the general case where both the meson partner and baryon partners are available, the quark will form a meson or a baryon according to the following criteria [32]:

dB<dM rBM:formabaryon;otherwise:formameson. (7)

In the above, rBM is the new coalescence parameter, which controls the relative probability of a quark forming a baryon instead of forming a meson. Note that the same coalescence procedure is also applied to all antiquarks, and the above criteria are not needed when only the meson partner or baryon partners can be found for a parton. In the limit of rBM0, there would be no antibaryon formation at all while the minimum number of baryons would be formed as a result of the conservation of the (positive) net-baryon number. On the other hand, in the limit of rBM, there would be almost no meson formation; more specifically, only 0, 1, or 2 mesons would be formed depending on the remainder when dividing the total quark number in the event by three. As a result, the new quark coalescence allows a (anti)quark the freedom to form a meson or a (anti)baryon depending on the distance from the coalescence partner(s). This is a more physical picture; for example, if a subvolume of the dense matter is only made of quarks which total number is a multiple of three, it would hadronize to only baryons (with no mesons) as one would expect.

We take central Au+Au collisions at sNN=200 GeV from the AMPT-SM model as an example to compare the old and new quark coalescence [32]. The same parton cross section is used so that the parton phase-space configuration just before quark coalescence is statistically the same for the old and new quark coalescence. Figure 4 shows the average coalescence time of partons in mesons and (anti)baryons as functions of the hadron rapidity yH. We see that baryons and antibaryons in the new quark coalescence (curve with open circles) are now formed much earlier than before. This is because the old quark coalescence tends to form (anti)baryons late, since it searches for meson partners before (anti)baryon partners and a parton will be unavailable for (anti)baryon formation when it is already used for meson formation. In contrast, the new quark coalescence searches for the potential meson partner and (anti)baryon partners concurrently and then determines the hadron type to be formed, making the coalescence process more physical as well as more efficient. In addition, we see that mesons in the new quark coalescence (curve with filled circles) are also formed earlier than before, possibly because of the improved efficiency after giving partons the freedom to form either a meson or a (anti)baryon. Since the plotted coalescence time is in the center-of-mass frame of the AA collision, we would expect a cosh yH dependence if the dense matter were boost-invariant. The dot-dashed curve in Fig. 4 represents a function that is proportional to cosh yH, which qualitatively agrees with our model results.

Fig. 4
(Color online) The average coalescence time of partons in mesons and (anti)baryons as functions of the hadron rapidity from the new (curves with circles) and old (dashed curves) quark coalescence for central Au+Au collisions at 200 GeV; the dot-dashed curve represents a cosh yH curve for comparison.
pic

Therefore, the new quark coalescence is more efficient, especially for the formation of (anti)baryons, due to the freedom of a parton to form either a meson or a (anti)baryon. As a result, it leads to improvements in the descriptions of (anti)baryon observables from the AMPT-SM model [32, 33, 113]. Figure 5 shows the AMPT results (with rBM=0.61) on various antiparticle-to-particle ratios around mid-rapidity for central Au+Au collisions at 200 GeV [38, 114, 115] and Pb+Pb collisions at 2.76 TeV [101, 116, 117] in comparison with the experimental data at mid-rapidity. Both the data and model results here are for the 0-5% centrality except that Ω at 200 GeV corresponds to the 0-10% centrality. We see that the results from the new quark coalescence (solid curves) are generally consistent with the experimental data, while results from the old quark coalescence (dashed curves) severely overestimate the ratios for Ξ and Ω. In addition, the antibaryon-to-baryon ratios generally increase with the strangeness content in both the AMPT model and the data. This is consistent with models such as the ALCOR model [118], which predict that these ratios are sequentially higher by a multiplicative factor, the K+/K- ratio. Since the K+/K- ratio is usually slightly larger than one at high energies, we see that our results from the improved quark coalescence agree rather well with this expectation and with the experimental data.

Fig. 5
(Color online) Antiparticle-to-particle ratios around mid-rapidity for (a) central Au+Au collisions at 200 GeV and (b) central Pb+Pb collisions at 2.76 TeV from the new (solid curves) and old (dashed curves) quark coalescence in comparison with the experimental data.
pic

On the other hand, the AMPT model with the improved quark coalescence [32, 119] still underestimates the p¯/p ratio in central Au+Au collisions at and below 200 GeV. We note that quark coalescence should be augmented with other hadronization mechanisms such as fragmentation [110, 111] for partons that cannot find nearby partners. This will also help avoid the potential violation of the second law of thermodynamics during the hadronization process [120], where whether the entropy decreases during a phase-space quark coalescence has been found to depend on details such as the duration of the mixed phase, volume expansion, and resonance decays [121]. Also note that the rBM value of 0.61 is found to reasonably reproduce the proton and antiproton yields of AA collisions in the AMPT model with the original parton distribution function and HIJING’s nuclear shadowing [32], while the preferred rBM value is 0.53 for light (u/d/s) hadrons [119, 122] and 1.0 for charm hadrons [122] in the AMPT model with modern parton distribution functions of nuclei.

The new quark coalescence is not only able to describe the dN/dy yields, pT spectra, and elliptic flows of pions and kaons at low pT, it also better describes the baryon observables in general, especially the pT spectra of (anti)baryons and antibaryon-to-baryon ratios for Ξ and Ω. It has also been shown to qualitatively describe the near-side anticorrelation feature of baryon-baryon azimuthal correlations observed in small systems at the LHC [33, 113]. In addition, it can be easily extended to include individual rBM factors specific to given hadron species, e.g., to describe the enhanced multi-strange baryon productions in nuclear collisions [123]. The string melting AMPT model with the new quark coalescence thus provides a better overall description of the bulk matter in high-energy nuclear collisions.

2.4  Importance of finite nuclear thickness at lower energies

For heavy ion collisions at lower energies, the thickness of the incoming projectile and target nuclei in the center-of-mass frame becomes larger due to the finite Lorentz contraction along the beam directions. Therefore, one needs to consider the finite nuclear thickness in dynamical models of heavy ion collisions at lower energies, which correspond to higher net-baryon densities. The finite nuclear thickness increases the longitudinal width of the created matter and thus will obviously affect the initial energy and net-baryon densities [124, 125]. Furthermore, it will lead to a significant time duration of the initial particle and energy production; therefore, one cannot use a fixed proper time to describe the initial condition for hydrodynamic-based models but use a dynamical initialization scheme [126, 127].

For a central collision of two identical nuclei of mass number A, it takes the following time for the two nuclei to completely cross each other in the center-of-mass frame:

dt=2RAsinhycm (8)

in the hard sphere model of the nucleus. In the above, RA is the hard-sphere radius of the nucleus and ycm is the rapidity of the projectile nucleus. For central Au+Au collisions at sNN=50 GeV, for example, dt≈0.5 fm/c, which is comparable to the typical value of the parton formation time or hydrodynamics initial time when one takes RA=1.12 A1/3 fm. Therefore, one may expect the effect from finite nuclear thickness to be significant for central Au+Au collisions at sNN50 GeV, which is the focus energy range of the RHIC Beam Energy Scan (BES) program.

We have developed semi-analytical methods [124, 125] to include the finite nuclear thickness in the calculation of the initial energy density, which is crucial in determining the initial temperature (and net-baryon chemical potential at low energies) of the produced QGP. Traditionally, the Bjorken formula [128] has been the standard semi-analytical tool in estimating the initial energy density in the central rapidity region right after the two nuclei pass each other:

ϵBj(t)=1AT tdETdy. (9)

In the above, AT represents the full transverse area of the overlap volume, and dET/dy is the initial rapidity density of the transverse energy at mid-rapidity, which is often approximated with the experimental dET/dy value in the final state. Because the Bjorken energy density diverges as t0, a finite value is needed for the initial time, which is often taken as the proper formation time of the produced quanta τf. However, a serious limitation of the Bjorken formula results from the fact that it neglects the finite thickness of the colliding nuclei. Therefore, one expects that the Bjorken formula may break down when the crossing time is not small compared to the formation time [6].

Using the semi-analytical methods that include the finite nuclear thickness, the initial energy density ϵ(t) averaged over the transverse area of the overlap region as a function of time, including its maximum value ϵmax, have been calculated [124, 125]. We first considered the finite time duration of the initial energy production but neglected the finite longitudinal extension [124], which enabled us to obtain explicit analytical solutions of ϵ(t). Both the uniform time profile and beta time profile have been considered, where in the uniform time profile one assumes that the initial transverse energy at y ≈ 0 is produced uniformly in time (x) from t2 to t2:

d2ETdydx=1t2t1dETdy, if x[t1,t2]. (10)

In contrast, the beta time profile assumes the following:

d2ETdydx[x(dtx)]ndETdy, if x[0,dt]. (11)

Note that n=4 is chosen [124] from the comparison to the time profile of partons within mid-spacetime-rapidity in central Au+Au collisions from the AMPT-SM model. In addition, for the uniform profile shown here, t1=0.29dt & t2=0.71 dt are used since they give the same mean and standard deviation of time as the beta profile at n=4.

Then we considered both the finite time duration and longitudinal extension of the initial energy production [125]. When τf is not too much smaller than the crossing time of the two nuclei, results from the latter study [125] are similar to those from the former study [124]. On the other hand, there is a qualitative difference in that the maximum energy density ϵmax at τf=0 is finite after considering both the finite duration time and longitudinal extension [125], while the Bjorken formula diverges as 1/τf and the method that only considered the finite duration time [124] diverges as ln (1/τf) at low energies but as 1/τf at high energies. Overall, these studies have yielded the following qualitative conclusions: the initial energy density after considering the finite nuclear thickness approaches the Bjorken formula at high colliding energies and/or large formation time τf. At low colliding energies, however, the initial energy density has a much lower maximum, evolves much longer, and is much less sensitive to τf than the Bjorken formula. Note that we have written a web interface [129] that performs the semi-analytical calculation [125] for central AA collisions, where the user can input the colliding system, energy and the proper formation time.

To include the finite nuclear thickness, we have modified the initial condition of the AMPT-SM model [124] to specify in each heavy ion event the longitudinal coordinate z0 and time t0 of each excited string, which is then converted into the initial partons via string melting. Note that in the normal AMPT-SM model [26, 13, 27, 32] the longitudinal coordinate z0 and time t0 of each excited string in the initial state are both set to zero, which would be correct only at very high energies. Figure 6 shows the results on the time evolution of the average energy density at ηs ≈ 0 for central Au+Au collisions at four different energies for formation time τf=0.1 and 0.3 fm/c. At the high energy of 200 GeV, the AMPT-SM results with (curves with filled circles) and without (curves with open circles) the finite nuclear thickness are essentially the same. This is consistent with the fact that the Bjorken result and our semi-analytical result are also very similar (after shifting the results in time); it also confirms the expectation that the finite nuclear thickness can be mostly neglected at high-enough energies.

Fig. 6
(Color online) Average parton energy densities at central spacetime-rapidity from AMPT with (filled circles) and without (open circles) the finite nuclear width for central Au+Au collisions and τf=0.1 & 0.3 fm/c at (a) 4.84, (b) 11.5, (c) 27, and (d) 200 GeV; corresponding results for the uniform time profile (dashed curves), the beta time profile (solid curves), and the Bjorken formula (dot-dashed curves) are also plotted for comparison.
pic

At lower energies, however, the AMPT results after including the finite nuclear thickness are very different: the maximum energy density is lower, and the time evolution of the energy density (e.g., the time spent above half the maximum energy density) is longer. These key features agree with the semi-analytical results [124, 125], where the results from the uniform time profile and the more realistic beta time profile are close to each other after the uniform profile is set to the same mean and standard deviation of time as the beta profile [124]. We also see from Fig. 6 that the increase of the maximum initial energy density with the colliding energy is much faster after including the finite nuclear thickness, which is consistent with the analytical finding that the Bjorken formula overestimates the maximum energy density more at lower energies [124, 125]. In addition, we see in Fig. 6 that the AMPT results are generally wider in time; partly because the parton proper formation time in AMPT is not set as a constant but is inversely proportional to the parent hadron transverse mass [13]. Secondary parton scatterings and the transverse expansion of the dense matter in AMPT can also cause differences from the semi-analytical results, which do not consider such effects. Overall, we see that the AMPT results without considering the finite nuclear thickness are similar to the Bjorken results, while the AMPT results including the finite thickness are similar to our semi-analytical results. These results suggest that it is important to include the finite nuclear thickness in dynamical models of relativistic heavy ion collisions, especially at lower energies.

2.5  Modern parton distribution functions in nuclei

The initial condition of the AMPT model is based on the HIJING two-component model [23]. The primary interactions between the two incoming nuclei are divided into two parts: the soft component described by the Lund string fragmentation model [25, 92, 93], and the hard component with minijet productions described by perturbative QCD through the PYTHIA5 program [25].

The minijet differential cross sections in HIJING model can be computed using the factorization theorem in the perturbative QCD framework [130] as

dσjetcddpT2dy1dy2=Ka,bx1fa(x1,Q2)x2fb(x2,Q2)dσabcddt^. (12)

In the above, pT is the transverse momentum of the produced minijet parton, y1 and y2 are the rapidities of the two produced partons c and d, the factor K accounts for higher-order corrections to the leading order jet cross section, x1 and x2 are respectively the momentum fraction x carried by the two initial partons, fa(x1, Q2) is the parton distribution function (PDF) of parton type a at x=x1 and factorization scale Q2, σab is the parton-parton cross section for partons a and b, and t^ is the standard Mandelstam variable for the minijet production subprocess.

The total inclusive jet cross section (for the production of minijet gluons and u/d/s quarks and antiquarks) is then obtained by integrating the above differential cross section with a transverse momentum cutoff p0 and considering all the possible combinations of final state parton flavors [23]:

σjet=12c,dp02s^/4dpT2dy1dy2dσjetcddpT2dy1dy2, (13)

where s^ is the standard Mandelstam variable for the minijet subprocess. We see that the minijet transverse momentum cutoff p0 and the parton distribution function fa(x,Q2) are the key factors affecting the jet cross section. The total jet cross section together with the σsoft parameter that describes the soft component determine the nucleon-nucleon interaction cross sections in the Eikonal formalism [131, 132]. Note that p0 is only relevant when the center-of-mass energy per nucleon pair is higher than 10 GeV, because the jet production in the HIJING model is switched off at sNN<10 GeV.

An important ingredient needed in Monte Carlo event generators for hadron collisions is the input parton distribution function [133-135]. Efforts have been made to understand the subtilties of implementing various parton distributions for phenomenological studies based on event generators[136, 137]. The impacts of different parton distributions in the event generators for pp collisions are found to be sizable and the key parameters in the generators usually depend on the details of the input PDF [138]. Specifically, the parton distribution function in the AMPT model affects the initial state radiation and the minijet production within the two-component model framework. Using modern parton distributions along with fine tuned model parameters is required to generate reliable exclusive final states in the AMPT model.

The HIJING 1.0 model [23, 24] that generates the initial condition of the original AMPT model employs the Duke-Owens parton distribution function set 1 [139] for the free proton. However, it is well known that the Duke-Owens PDFs were obtained at a time when a large array of experimental data used in the global fittings for modern PDFs were not yet available [135]. The parton densities at small-x relevant for minijet and heavy flavor productions at high energies are generally underestimated by the Duke-Owens PDFs [140]. Therefore, we have updated the AMPT model with a modern parton distribution function of the free nucleon (the CTEQ6.1M set [141]) and retuning of the relevant p0 and σsoft parameters [119]. Note that this update is based on the AMPT model with the new quark coalescence [32]. Also note that the HIJING 2.0 model [142] is a similar update, which replaces the Duke-Owens PDFs in the HIJING 1.0 model with the GRV PDFs [143].

For nuclear collisions at sufficiently high energies, results from event generators depend on the parton distribution functions inside the incoming nuclei. Analogous to the free nucleon case, global analyses of the parameterizations for the modifications of the nuclear PDFs relative to the free nucleon PDFs have been performed by several groups [144-148]. In addition, it is natural to expect the nuclear modification to depend on a nucleon’s position inside a nucleus. Therefore, the spatial dependence of nuclear parton densities are considered [23, 149-154], and a global analysis to extract the nuclear PDFs with spatial dependence is carried out [155] based on the EKS98 [156] and EPS09 [157] fits. In a recent study [119], we have included the spatially dependent EPS09s nuclear modifications [155] in the AMPT model to replace the original HIJING 1.0 nuclear shadowing parameterization. Note that the HIJING 1.0 shadowing is spatially dependent but independent of Q2 or the parton flavor [13, 23], similar to the HIJING 2.0 nuclear shadowing [142].

For a proton bound in a nucleus, its parton distribution function of flavor i can be written as

fip/A(x,Q2)RiA(x,Q2)fip(x,Q2), (14)

where fip(x,Q2) is the corresponding PDF in the free proton. Here RiA(x,Q2) represents the spatially averaged nuclear modification function, which typically depends on the x range: the shadowing region at small x, an anti-shadowing region at x ˜0.1, and the EMC effect region at x close to 1. The spatial dependence of the nuclear modification function can be formulated as

RiA(x,Q2)1Ad2s TA(s) riA(x,Q2,s), (15)

where TA(s) represents the nuclear thickness function at transverse position s, and riA(x,Q2,s) represents the spatially dependent nuclear modification function.

Solid curves in Fig. 7 show the gluon density distributions (multiplied by x) in the free proton from the original and the updated AMPT model. The gluon density distributions of a bound proton at the center of a lead nucleus from the EPS09s and HIJING nuclear modifications are also shown in Fig. 7. We see that in the updated AMPT model with the CTEQ6.1M set the gluon densities are quite different from the old Duke-Owens set and much higher at small x. We also see that the gluon shadowing in EPS09s is much weaker than that in the HIJING 1.0 model.

Fig. 7
(Color online) Gluon density distributions (multiplied by x) in free proton (solid curves) and proton inside the lead nucleus (dashed curves) at Q2=10 GeV2 from the original (black) and updated (red) AMPT.
pic

As mentioned earlier, p0 and σsoft are the two key parameters in the HIJING model that determine the total and inelastic cross sections of pp and pp¯ collisions within the Eikonal formalism. In the HIJING 1.0 model that uses the Duke-Owens PDFs, constant values of p0 = 2.0 GeV/c and σsoft = 57 mb are found to reasonably describe the experimental cross sections of pp and pp¯ collisions over a wide energy range [23, 24, 158]. On the other hand, when the PDFs are updated in the HIJING 2.0 model [119, 142], energy-dependent p0(s) and (s) values are needed.

In our work that implements the CTEQ6.1M PDF in the AMPT model [119], the energy dependent parameters p0(s) and (s) are determined via fitting the experimental total and inelastic cross sections of pp and pp¯ collisions within the energy range 4<s<105 GeV, as shown in Fig. 8. We then obtain the following p0(s) and (s) functions for pp collisions:

Fig. 8
(Color online) Total and elastic cross sections versus the colliding energy of pp collisions from the updated AMPT model (solid and dot-dashed curves) in comparison with the experimental data (symbols); σjet from the model is also shown (dotted curve).
pic
p0pp(s)=1.71+1.63ln(s)0.256ln2(s)+0.0167ln3(s), (16) σsoft(s)=45.1+0.718ln(s)+0.144ln2(s)+0.0185ln3(s). (17)

In the above, p0pp and the center-of-mass colliding energy s are in the unit of GeV/c and GeV, respectively; while σsoft is in the unit of mb. Note that s will be replaced with sNN for nuclear collisions. We also find that the updated AMPT-SM model reasonably describes the charged particle yield and pT spectrum in pp and/or pp¯ collisions from s 4 GeV to 13 TeV [119].

When we apply the above p0(s), (s) and the EPS09s nuclear shadowing to central AA collisions at LHC energies, however, we find rather surprisingly that the hadron yields from the AMPT-SM model are significantly higher than the experimental data. As shown in Fig. 9(a), the AMPT-SM model that uses the p0pp(s) value overestimates the final state hadron multiplicities in central Pb+Pb collisions at sNN=2.76 TeV. Since a larger p0 value would suppress the total jet cross section and reduce the particle yields, we introduce a global scaling of the minijet cutoff p0 to make its value in central AA collisions, p0AA(s), nuclear size dependent [119]:

Fig. 9
(Color online) Identified particle rapidity distributions in 0-5% central Pb+Pb collisions at sNN=2.76 TeV from the AMPT-SM model using the minijet cutoff (a) p0pp(s) or (b) p0AA(s) in comparison with the ALICE data [99, 101].
pic
p0AA(s)=p0pp(s)Aq(s),q(s)=0.0334ln(sE0)0.00232ln2(sE0)+0.0000541ln3(sE0),for sE0=200 GeV. (18)

The above q(s) function is determined from the fit to the overall particle yields of central Au+Au collisions at the RHIC energies and central Pb+Pb collisions at LHC energies (see [119] for details). Its value is zero at sNN200 GeV since the p0pp(s) value works reasonably well there for central Au+Au collisions, while its value approaches 0.16 at sNN107 GeV. This nuclear scaling of the minijet momentum cutoff scale p0 is motivated by the physics of the color glass condensate [159], where the saturation momentum scale Qs depends on the nuclear size as QsA1/6 in the saturation regime for small-x gluons in AA collisions at high-enough energies.

As shown in Fig. 9(b), the updated AMPT-SM model using p0AA(s) from the global nuclear scaling well reproduces the identified particle yields in central Pb+Pb collisions at the LHC energy. In addition, we find that a significantly smaller value for the Lund bL parameter, bL=0.15 GeV-2, is needed to describe the particle pT spectra in central AA collisions [119], similar to an earlier study [27]. Note that recently we have generalized the minijet cutoff p0 and the Lund bL parameter with a local nuclear scaling [160], as shall be discussed in Sec. 2.7, which would help explain why a bigger p0 value but a smaller bL value are needed for high energy AA collisions than pp collisions.

2.6  Improvements of heavy flavor productions

Heavy flavors are predominantly produced from the initial hard scatterings at early times in nuclear collisions [161-163]. Therefore, they are powerful observables to probe the strong electromagnetic field created in heavy ion collisions [164-166] and transport properties of the dense matter [167-171]. Multiple theoretical frameworks have been developed for the description of open heavy flavor productions in high energy pp and pA collisions based on the pQCD framework [172-175]. Medium effects, such as those from pQCD calculations of the parton energy loss [176, 177] or the Langevin/Boltzmann equation methods [178-186] can be included for AA collisions.

Study of heavy flavor productions within the AMPT model [187, 188] has the potential to provide a unified model for both light and heavy flavor transport and improve our understanding of the non-equilibrium effects of the QGP evolution [49, 50, 189, 190]. In addition, using parton scatterings to model the interactions between heavy quarks and the evolving medium, the parton cascade approach is able to implement any scattering angular distribution without the need to assume small-angle scatterings. Therefore, after the update with modern parton distributions for proton and nuclei as discussed in Sec. 2.5, we have made several significant improvements of heavy flavor productions in the AMPT model [122]. First, for self consistency we include the heavy flavor cross sections in the total minijet cross section in the HIJING two-component model. Second, we remove the minimum transverse momentum requirement (p0) for initial heavy quark productions since the heavy quark pair production cross sections from pQCD are already finite due to the heavy quark mass. These changes can be illustrated with the following modified formula for the minijet cross section [122]:

σjet=c,d11+δcdp02s^/4dpT2dy1dy2dσlightcddpT2dy1dy2+c,d0s^/4dpT2dy1dy2dσheavycddpT2dy1dy2, (19)

where the first term on the right hand side represents the cross section of light flavor (u/d/s/g) minijets and the second term represents that of heavy flavors such as charm and bottom. Note that the factor 1/(1+δcd) above becomes 1/2 for final states with identical partons, such as g+gg+g for minijet gluon productions. In contrast, the original HIJING model uses Eq.(13) and applies the factor of 1/2 to all light flavor minijet production processes [23], and that leads to a smaller σjet than Eq.(19) (at the same p0). As a result, an increase of the p0 value as shown below is needed [122] to describe the experimental data on total and inelastic cross sections of pp and pp¯ collisions shown in Fig. 8:

p0pp(s)=1.92+1.77ln(s)0.274ln2(s)+0.0176ln3(s) (20)

with p0 in GeV/c and s in GeV.

The total cc¯ cross section for pp collisions from the updated AMPT model (solid curve) is shown in Fig. 10 versus the colliding energy in comparison with the available world data. We see that the updated AMPT model can well describe the data in pp collisions at various collision energies. The original AMPT model (dotted curve), however, significantly underestimates the charm quark yield, especially at low energies. The enhanced charm quark production in the updated model is largely due to the removal of the p0 cut, since the charm quark cross section is much lower when charm quarks in the updated AMPT model are required to have a transverse momentum above p0 (dashed curve).

Fig. 10
(Color online) Total cross sections of charm-anticharm quark pairs from the AMPT model for pp collisions in comparison with the world data [191-197] as functions of the colliding energy.
pic

Figure 11 shows the charm quark rapidity and transverse momentum distributions from the AMPT model for pp collisions at s=200 GeV and 7 TeV. Note that the charm quark or hadron results shown in this section have been averaged over those for particles and the corresponding antiparticles. As shown in Fig. 11(a), the charm quark yield in the updated AMPT model is found to be significantly higher than that in the original AMPT model over the full rapidity range at both RHIC and LHC energies. From the charm quark pT spectra at mid-rapidity shown in Fig. 11(b), we see that the removal of the p0 cut for charm quarks mostly enhances charm quark productions at low pT. We also see that results from the original AMPT model (dotted curve) and the updated AMPT model that includes the p0 cut (dashed curve) are similar, partly because a p0 cut (2 GeV/c) on the charm quark production is also used in the original AMPT model.

Fig. 11
(Color online) Charm-anticharm quark pair (a) rapidity distributions and (b) pT spectra around mid-rapidity in pp collisions at s=200 GeV (red curves) and 7 TeV (black curves) from the AMPT model. The shaded band represents statistical errors.
pic

In AA collisions, heavy quark production is subject to additional medium-induced initial state and final state effects. Within the AMPT model, initial state effects include the nuclear modification of the parton distribution functions in nuclei, while the final state effects are mostly treated with parton elastic rescatterings in the parton cascade [109]. Figure 12 shows the charm quark yield at mid-rapidity for 0-10% central Au+Au or 0-10% central Pb+Pb collisions at different energies. We see that the EPS09s nuclear modification leads to an enhancement of the charm quark yield in central AA collisions at lower energies but a suppression at high energies. This is expected due to the anti-shadowing feature at large x and the shadowing feature at small x in the nuclear modification functions. We also see that the result from the updated AMPT model (solid curve) is in good agreement with the charm quark data, which is obtained for 0-10% central Au+Au collisions at sNN=200 GeV by scaling the STAR pp charm quark cross section data with the number of binary collisions [198, 199]. Similar to the results for pp collisions, the updated AMPT model gives a significantly higher charm quark yield at mid-rapidity in central AA collisions compared to the original AMPT model.

Fig. 12
(Color online) dN/dy of charm quark pairs around mid-rapidity from the AMPT model for 0-10% central Au+Au collisions at RHIC energies and Pb+Pb collisions above RHIC energies as functions of the colliding energy in comparison with the STAR data.
pic

The open heavy flavor hadron species formed by quark coalescence include charm and bottom hadrons with all possible charges. To reproduce the observed vector to pseudo-scalar meson ratios of open heavy flavors in pp collisions, we fit the relative probability of forming primordial vector versus pseudo-scalar heavy mesons in the quark coalescence model, e.g., the ratio is set to 1.0 for the primordial D*/D and B*/B ratios [122], instead of using only the invariant mass of the coalescing partons in the original AMPT-SM model [13]. Note that only the primordial ratios right after coalescence are specified with these parameters, not the vector to pseudo-scalar meson ratios in the final state which include effects from resonance decays. In addition, in the new quark coalescence model [32] that is used in this heavy flavor work [122], the relative probability for a quark to form a baryon instead of a meson is determined by the rBM parameter as shown in Eq.(7). In our earlier work that updated the AMPT model with modern PDFs [119], the rBM value for light flavor (u/d/s) hadrons is set to 0.53, which value is also used here. On the other hand, we set the rBM value for heavy flavor hadrons to 1.0, because using the light flavor value would lead to too few charm baryons (by a factor of ˜4) compared to the experimental data in pp or AA collisions. In principle, the rBM value for charm hadrons depends on properties such as the number and masses of available charm baryon states versus charm meson states. The necessity of using a higher rBM value for charm is consistent with the assumption that there are more charm baryon states than charm meson states compared to the light flavor sectors [200].

After the improvements on heavy flavor productions, we find that the updated AMPT model [119] can well describe the yields and pT spectra of open charm hadrons including D, D*, Ds and Λc in pp collisions at different energies. The updated model also describes the charm data in central AA collisions much better than the original AMPT model. However, the updated AMPT model still does not well describe the charm hadron productions in AA collisions [122]. As shown in Fig. 13(a), the updated AMPT model overestimates the D0 yield at low pT but underestimates it at pT above 2.5 GeV/c when compared to the STAR data for 0-10% Au+Au collisions at sNN=200 GeV [198]. Compared to the original AMPT model, the charm hadron yield from the updated AMPT model is significantly enhanced at low pT, similar to the results at the parton level shown in Fig. 11(b). We also find that the charm baryon to meson ratio (Λc/D) in 0-10% Au+Au collisions at 200 GeV is much larger in the updated AMPT model (dashed curve) than the original AMPT model (dotted curve), as shown in Fig. 13(b). Compared to the STAR data for 10-80% Au+Au collisions at 200 GeV [201], result of the Λc/D ratio versus pT from the updated AMPT model is somewhat higher. We also see that this ratio from the AMPT model is slightly lower in the 10-80% centrality than the 0-10% centrality for Au+Au collisions at 200 GeV.

Fig. 13
(Color online)(a) pT spectra of open charm hadrons in central Au+Au collisions at sNN=200 GeV and (b) the Λc/D ratios versus pT for two centralities of Au+Au collisions at 200 GeV from the AMPT model in comparison with the STAR data [198, 201].
pic

We note that only elastic parton scatterings are included in the AMPT-SM model; therefore, the model is only applicable in the region where the effect of parton radiative energy loss is small. Studies have suggested that the elastic collisional energy loss could be dominant for charm hadrons below pT ˜5-6 GeV/c in Au+Au collisions at sNN = 200 GeV or below pT ˜15 GeV/c in Pb+Pb collisions at sNN = 2.76TeV [184]. Therefore, the charm results from the AMPT model are not reliable at pT higher than these values. Also note that the charm pT spectra are affected by the charm quark scattering cross section and its angular distribution in ZPC. The AMPT model currently uses the g+gg+g cross section for scatterings of all parton flavors, where flavor-dependent cross sections and angular distributions should be used for the parton scatterings. In addition, there is still a large uncertainty on the nuclear shadowing of gluons [155], which has not been fully explored in the AMPT model. Furthermore, hadronic scatterings of heavy flavor hadrons [202-206] have not been included in the AMPT model except for the decays of heavy hadron resonances. Further developments of the AMPT model are expected to improve its description of heavy flavor productions in AA collisions.

2.7  System size dependence under local nuclear scaling

The system size dependence of observables can be useful to uncover the transition of certain phenomena in nuclear collisions, such as the onset of collectivity and whether it comes from initial state momentum correlations [54, 55] or final state interactions [49-51, 56, 57]. It is known from multiple studies that certain key parameters in the initial condition of the AMPT model for AA collisions need to be different from their values for pp collisions to reasonably describe the data [13, 22, 27, 32, 95, 119]. First, the Lund bL parameter in the symmetric string fragmentation function [92, 93], as shown in Eq.(5), for large collision systems needs to be significantly smaller than its value for pp collisions. An earlier study has also shown that a constant bL can not describe the centrality dependence of 〈pT〉 in heavy ion collisions [31], where the system size dependence of the Lund fragmentation parameters was suggested as a possible solution. Note that similar frameworks for the system size dependence have been implemented in the string fragmentation model [207-211]. Second, we have found in earlier developments of the AMPT model [119, 122] that the minijet transverse momentum cutoff p0 for central Pb+Pb collisions at the LHC energies needs to be significantly higher than that for pp collisions at the same energy. These observations suggest that the above two parameters should be related to the size of the colliding system to provide better initial conditions for the AMPT model.

Therefore, we have recently proposed [160] that the bL and p0 parameters in AMPT can be considered as local variables that depend on the nuclear thickness functions of the two incoming nuclei. This prescription allows us to use the parameter values obtained for pp collisions and the local nuclear scaling relations to obtain the values for AA collisions; the model would then describe the system size and centrality dependences of nuclear collisions self-consistently.

In the Lund string model [92, 93], the symmetric fragmentation function is given by Eq.(5). The average squared transverse momentum of massless hadrons is related to the Lund fragmentation parameters aL and bL as [13]

pT2=1bL(2+aL). (21)

Consequently, the 〈pT〉of partons after string melting and the final hadrons are significantly affected by the value of bL. Since the mean transverse momentum of initial partons in heavy ion collisions is expected to be higher in larger systems due to the higher initial temperature, we expect the bL value to decrease with the system size. Note that the string tension is believed to be larger in a denser matter [209, 210, 212, 213], thus a decrease of bL with the system size is consistent with the expectation of a stronger color field and thus a higher string tension κ since κα/ 1=bL[13] as shown in Eq.(6).

bL(sA,sB,s)=bLpp[TA(sA)TB(sB)/Tp]β(s). (22)

In the above, bLpp is the value for pp collisions (chosen to be 0.7 GeV-2 based on the fit of 〈pT〉 data), s represents the square of the center-of-mass collision energy per nucleon pair, TA(sA)=ρA(sA,z)dz is the nuclear thickness function at the transverse distance sA from the center of nucleus A determined with the Woods-Saxon nuclear density profiles [156], and Tp is the average value of the effective thickness function of the proton (taken as 0.22 fm-2). Note that Tp is used instead of TA(sA) or TB(sB) when the projectile or the target is proton or when TA(sA) or TB(sB) from the nucleus is smaller than the Tp value. Although different mathematical forms from that of Eq.(22) can be used in the local scaling relation, our study [160] shows that the geometric scaling functions (i.e., using the geometric mean of the two nuclear thickness functions) generally works better than the arithmetic form. We note that a systematic Bayesian analysis based on the TRENTo initial condition [214] with a hybrid model found that the geometric form for the local nuclear scaling is preferred by the experimental data [215].

The exponent function β(s) describes the energy dependence of the local nuclear scaling of bL. From the fits to charged particle 〈pT〉 data in the most central Au+Au collisions at RHIC energies and most central Pb+Pb collisions at LHC energies, it is parameterized as

β(s)=0.620+0.112ln(sE0)Θ(sE0), (23)

where E0=200 GeV and Θ(x) is the unit step function. The fitted β(s) function is shown in Fig. 14(a) (dashed curve), which is a constant at RHIC energies but grows rapidly at LHC energies. Note that β=1 at high energies (dotted line) may be a “natural” limit for Eq.(22) if we imagine that all local strings fully overlap so that the string tension adds up. That would give bL1/TA(sA) for central AA collisions, where TA(sA) is proportional to the local number of participant nucleons or excited strings integrated over the longitudinal length.

Fig. 14
(Color online)(a) Fitted exponent functions α(s), β(s), and 3q(s) versus the center-of-mass energy per nucleon pair s. (b) Average p0 and bL values versus the impact parameter of pPb, Au+Au and Pb+Pb collisions at several energies.
pic

Figure 14(b) shows the bL value averaged over the overlap volume versus the impact parameter for Pb+Pb and pPb collisions at 5.02A TeV and Au+Au collisions at two RHIC energies. We see that 〈bL〉 for Pb+Pb collisions at the LHC energy is lower than that for Au+Au collisions at RHIC energies, which corresponds to a larger string tension due to the larger value of the exponent β(s) at LHC energies. On the other hand, the impact parameter dependences of 〈bL〉 at different RHIC energies are essentially the same since β(s) is a constant within that energy range. For pPb collisions at 5.02A TeV, its 〈bL〉 is higher than that in Pb+Pb collisions at small b and grows faster with b due to its smaller system size.

The minimum transverse momentum cutoff p0 for light flavor minijet productions is another key parameter in the HIJING model and thus in the initial condition of the AMPT model [23, 119, 142]. In our update of the AMPT model with modern nPDFs [119], the collision energy dependence of p0 is determined from fitting the pp cross section data. Then motivated by the physics of the color glass condensate [159], a global nuclear scaling of the p0 cutoff [119] has been introduced for central AA collisions above the top RHIC energy of 200A GeV to describe the experimental data on charged particle yields in central Pb+Pb collisions at LHC energies. Here [160] we go beyond the global nuclear scaling and instead consider p0 as a local variable that depends on the transverse position of the corresponding hard process in each event. As p0 is expected to increase with the system size, we related its value to the nuclear thickness functions in a general AB collision as [160]

p0(sA,sB,s)=p0pp(s)[TA(sA)TB(sB)/Tp]α(s). (24)

As TA(s)A1/3, Eq.(23) approximately gives p0Aα(s)/3 for central AA collisions and thus essentially recovers the global nuclear scaling if α(s)=3q(s). On the other hand, for peripheral collisions where TA(sA) and TB(sB) are very small and thus replaced with the proton value (Tp), Eq.(23) automatically gives the p0 value for pp collisions.

Since p0pp(s) works for charged particle yields in central Au+Au collisions at and below 200A GeV, we assume that the need to modify p0 in nuclear collisions starts at the top RHIC energy [119]. From the comparison to charged particle yields in the most central Pb+Pb collisions at 2.76A and 5.02A TeV, we obtain the preferred α(s) values at those two energies. We then fit the α(s) function as [160]

α(s)=0.0918ln(sE0)0.00602ln2(sE0)+0.000134ln3(sE0), for sE0, (25)

with α(s)=0 for s<E0=200 GeV. We see in Fig. 14(a) that α(s) ≈ 3q(s) as expected. We also see that both approach the value of 1/2 at very high energies; this is consistent with our expectation [119] that p0 is closely related to the saturation momentum Qs in the color glass condensate [159], where QsA1/6 in the saturation regime.

Figure 14(b) also shows the average p0 value as a function of the impact parameter for Pb+Pb collisions at 2.76A TeV and 5.02A TeV as well as pPb collisions at 5.02A TeV. As expected, we see that 〈p0〉 decreases with the impact parameter and that 〈p0〉 at the lower LHC energy is smaller than that at the higher LHC energy due to the smaller α(s) value. Also, 〈p0〉 in pPb collisions is smaller than that in Pb+Pb collisions at the same colliding energy due to its smaller size. In addition, the relative variation of 〈p0〉 with the impact parameter is seen to be much weaker than that of 〈bL〉 since α(s)β(s) for the exponents in the local nuclear scaling relations.

We show in Fig. 15 the dNch/dη yield in panel (a) and charged particle 〈pT〉 in panel (b) around mid-pseudorapidity versus centrality from different AMPT versions in comparison with the experimental data for Au+Au collisions at 200A GeV and Pb+Pb collisions at 5.02A TeV [216-222]. Using the local nuclear scaling, the improved AMPT model (solid curves) reasonably describes these centrality dependence data in AA collisions at both RHIC and the LHC energies, with a significant improvement in the 〈pT〉 description as shown in Fig. 15(b). When we switch off the local nuclear scaling of p0 and bL but instead use constant bL=0.15 GeV-2 and p0(s) (constant at a given energy), we recover the AMPT model developed earlier [122] and obtain the dot-dashed curves when using p0(s)=p0AA(s) and the dotted curves when using p0(s)=p0pp(s). They both give the wrong centrality dependence of 〈pT〉, since the model results (dot-dashed or dotted) show a mostly increasing trend from central to peripheral collisions while the data show a mostly decreasing trend. Results from the public AMPT version 2.26t9 [61] are also shown (dashed curves) [27], which also fail to describe the centrality dependence of charged particle 〈pT〉 data.

Fig. 15
(Color online)(a) dNch/dη and (b) 〈pT〉 around mid-pseudorapidity versus the centrality of 5.02A TeV Pb+Pb collisions (thick curves) and 200A GeV Au+Au collisions (thin curves) from this work (solid curves) and earlier AMPT versions in comparison with the experimental data (symbols). The pT range used for the 〈pT〉 calculation is [0.15, 2] GeV/c at 5.02A TeV and [0.2, 2] GeV/c at 200A GeV.
pic

In Fig. 15(a), we see that the charged particle yield in central Pb+Pb collisions at 5.02A TeV is significantly overestimated when using p0(s)=p0pp(s), and the global nuclear scaling p0(s)=p0AA(s) is needed to reproduce the particle yield. From the Pb+Pb results, we also see that the effect from the global nuclear scaling of p0 in peripheral collisions is much smaller than that in central collisions, because the binary scaling of minijet productions makes p0 less important for peripheral collisions. It is thus not surprising to see that the dNch/dη results from the local nuclear scaling are similar to the AMPT results using the constant p0AA for central collisions but close to the AMPT results using the constant p0pp for peripheral collisions.

The local nuclear scaling relations also predict how observables depend on the system size going from large to small systems. Figures 16(a) and 16(b) show respectively the dNch/dη and charged particle 〈pT〉 around mid-pseudorapidity from the AMPT model [160] versus centrality in comparison with the experimental data for Au+Au collisions and several smaller collision systems [221, 223-226]. We see that the improved AMPT model describes these data rather well, further demonstrating the validity of the local nuclear scaling assumption. Note that, although the mid-pseudorapidity dNch/dη and 〈pT〉 data for the most central Au+Au/Pb+Pb collisions have been used in the determination of the parameter functions α(s) and β(s), the data of these smaller systems are not considered in the fitting of the parameters. Results for Au+Au collisions at 200A GeV are also presented in Fig. 16, and we see that the changes of the charge particle yield and 〈pT〉 from Cu+Cu to Au+Au collisions at 200A GeV are well accounted for by the local nuclear scaling. For example, the 〈pT〉 in Cu+Cu is generally smaller than that in Au+Au due to the larger bL value for Cu+Cu collisions. Note however that our calculations here have not considered the deformation of the Xe nucleus [227].

Fig. 16
(Color online) (a) dNch/dη and (b) 〈pT〉 around mid-pseudorapidity versus the centrality of Xe+Xe collisions at 5.44A TeV, Cu+Cu collisions at 200A GeV, and pPb collisions at 5.02A TeV from the AMPT model (curves) in comparison with the experimental data (symbols). The pT range used for the 〈pT〉 calculation is [0.15, 2] GeV/c at LHC energies and [0.2, 2] GeV/c at 200A GeV.
pic
2.8  PYTHIA8 initial condition with sub-nucleon structure

The modifications of the AMPT initial condition discussed so far have been performed within the framework of the HIJING two-component model that uses the PYTHIA5 program. While the development of local nuclear scaling [160] enables the AMPT model to reproduce the system size dependence and centrality dependence of changed particle yields and 〈pT〉 in pA and AA collisions using the parameter values for minimum bias pp collisions, we have not directly addressed the multiplicity dependence of these observables, especially the 〈pT〉, in pp collisions. On the other hand, PYTHIA8 [228] is quite successful in describing the particle yields in pp collisions. It has been extended to treat pA or AA collisions based on the Angantyr framework [229], and PYTHIA8 has been used as the initial condition generator for multiple heavy ion Monte Carlo models [230-232]. Therefore, it is worthwhile to have the option to use PYTHIA8 as the initial condition for the AMPT model.

Recently we have coupled PYTHIA8 with the final state parton and hadron interactions and quark coalescence [32] of the AMPT-SM model to study pp collisions [233]. In this approach, the fluctuating initial condition of AMPT originally provided by the HIJING model is replaced by the PYTHIA/Angantyr model [229]. In addition, the sub-nucleon structure, which could be important for collectivity observables in small systems [234-238], can be modeled when implementing the space-time structure of the string system generated by the PYTHIA multiparton interaction. With the proton charge distribution given by

ρ(r)=18πR3er/R (26)

with R=0.2 fm, the sub-nucleon spatial structure can be related to the transverse positions of the excited strings in two ways. In the first way, the transverse coordinates of the produced string objects are sampled according to the overlap function of a pp collision at a given impact parameter b:

T(x,y,b)=ρ(xb/2,y,z)ρ(x+b/2,y,z)dz, (27)

where z is along the beam directions. In the second way, the initial transverse spatial condition including event-by-event sub-nucleon fluctuations is generated with a Glauber Monte Carlo method based on the constituent quark picture [236, 239-243]. By modeling the proton as three constituent quarks, the interaction of two protons can be interpreted as collisions between the constituent quarks from each incoming proton within the Glauber model framework [241, 244]. The positions of the quark constituents are first sampled with the proton profile ρ(r), then the transverse coordinates of the excited strings are randomly assigned to the binary collision center of each interacting constituent pair.

Figure 17 shows the effect of using PYTHIA8 as the AMPT initial condition on the identified particle 〈pT〉 versus the charge particle pseudo-rapidity density in pp collisions at s=13 TeV. Note that only hadrons within 0<pT<3 GeV/c and |y|<0.5 are included in this comparison, and the central values of ALICE data are obtained with a refit to the data [245]. We see that this AMPT model (solid curves), which uses the PYTHIA8 initial condition and includes both parton and hadron evolutions, roughly reproduces the experimental data. On the other hand, the original AMPT model (dashed curves) reasonably describes the pion 〈pT〉 but gives a very weak multiplicity dependence for the proton 〈pT〉. The significant improvement compared to the original AMPT model on the multiplicity dependence of the proton 〈pT〉 presumably results from multiparton interaction in the PYTHIA8 model.

Fig. 17
(Color online) 〈pT〉 of π (black), K (blue) and proton (red) at mid-rapidity within 0<pT<3 GeV/c versus the charged hadron multiplicity density in 13 TeV pp collisions. The AMPT model using the PYTHIA8 initial condition (solid curves) are compared to the original AMPT model (dashed curves) and the ALICE data.
pic

Figure 18(a) shows the average initial spatial eccentricity of partons in the transverse plane right after string melting as a function of the parton multiplicity of each event from the two ways of generating the sub-nucleon spatial structure. Note that only partons with formation time less than 5 fm/c are considered, and eccentricities are calculated with the initial position of each parton at its formation time [246]. When using the overlap function weighting method (black curves), the eccentricity is largely driven by the geometric shape of the transverse overlap area and thus decreases significantly with the parton multiplicity as shown in panel (a) and increases significantly with the impact parameter as shown in panel (b). On the other hand, when using the Monte Carlo method with constituent quarks (red curves), large eccentricities in the initial condition can be generated even in very central collisions or events at high multiplicities. Figure 18(b) actually shows that the initial eccentricity from the constituent quark method is larger for pp collisions at smaller impact parameters, opposite to the behavior from the overlap function method.

Fig. 18
(Color online) The initial eccentricity of partons right after string melting versus (a) the number of partons and (b) the impact parameter, and (c) two-particle long-range angular correlations for events at two different multiplicity classes, for pp collisions at 13 TeV. AMPT results with the sub-nucleon structure are shown for the overlap function method (black curves) and the MC method with constituent quarks (red curves).
pic

The difference in the initial spatial eccentricity could certainly affect final state momentum anisotropies in small collision systems after interactions in the AMPT model convert the spatial anisotropies into momentum anisotropies [26, 49, 50]. Using the AMPT model with PYTHIA8 as the initial condition, we have found [233] that two-particle long-range correlations in high multiplicity pp collisions at the LHC depend sensitively on how the sub-nucleon structure of the proton is implemented. We analyze the projected correlation function of two charged hadrons with a large pseudorapidity gap:

C(Δϕ)=1NtrigdNpairdΔϕ. (28)

Both trigger and associate hadrons are required to be within 1<pT<3 GeV/c and |η|<2.4 following the analysis procedure of the CMS Collaboration [247], and the two hadrons in each pair must be separated in pseudo-rapidity with a gap |Δη|>2. Events are separated into two categories based on Nsel, the number of selected charge tracks with pT>0.4 GeV/c and |η|<2.4. High multiplicity events are defined as those with Nsel>80, while low multiplicity events are defined as those with Nsel<20.

Figure 18(c) shows the multiplicity dependence of the Cϕ) function from the two ways of generating the sub-nucleon spatial structure for 0.2 mb parton cross section [233]. We see that the AMPT model using PYTHIA8 shows a long-range ridge-like structure for high multiplicity events when the proton geometry is modeled with the constituent quark method (red solid curve), while the overlap function weighting method (black solid curve) does not show this structure. This demonstrates the connection between two-particle long-range correlations with the underlying sub-nucleon structure and fluctuations. Note that a significant near-side ridge structure in the correlation function is found in the experimental data, which has been regarded as an important signature of collectivity in high multiplicity pp events [44, 247].

We note that the original AMPT-SM model also shows the long-range near-side correlations, although it does not include the sub-nucleon structure [233]. In addition, the PYTHIA event generator itself has considered final state hadronic rescatterings [206, 248-250]. Using the AMPT-SM model with PYTHIA8 initial conditions, we can extend the study of pp collisions [233] to pA and AA collisions with the Angantyr model within the PYTHIA8 framework. That would lay a solid foundation for the studies of different mechanisms of collectivity, such as string shoving and parton/hadron evolutions, with the same model.

2.9  Improved algorithm for the parton cascade

Particle correlations and momentum anisotropies in the AMPT-SM model are usually dominated by parton interactions [13, 26, 41]. We have also found that even a few parton scatterings in a small system is enough to generate significant momentum anisotropies through the parton escape mechanism [49, 50]. It is therefore important to ensure that the parton cascade solution in the AMPT model is accurate.

The ZPC elastic parton cascade [109] in the AMPT model solves the Boltzmann equation by the cascade method, where a scattering happens when the closest distance between two partons is less than the range of interaction σp/π with σp being the parton scattering cross section. The default differential cross section in ZPC for two-parton scatterings, based on the gluon elastic scattering cross section as calculated with QCD at leading order, is given by [13, 109]

dσpdt^=9παs22(1+μ2s^)1(t^μ2)2, (29)

where μ is a screening mass to regular the total cross section. This way the total cross section has no explicit dependence on s^:

σp=9παs22μ2. (30)

The above Eqs.(28-29) represent forward-angle scatterings. For isotropic scatterings, dσp/dt^ is independent of the scattering angle.

It is well known that cascade calculations suffer from the causality violation [251, 252] due to the geometrical interpretation of cross section. This leads to inaccurate numerical results at high densities and/or large scattering cross sections (i.e., large opacities). For example, a recent study [29] has shown that the effect of causality violation on the elliptic flow from the AMPT-SM model [13] is small but non-zero. Causality violation also leads to the fact that different choices of performing collisions and/or the reference frame can lead to different numerical results [253-255]. These numerical artifacts due to the causality violation can be reduced or removed by the parton subdivision method [12, 43, 252, 254, 256260]. However, parton subdivision usually alters the event-by-event correlations and fluctuations, the importance of which has been more appreciated in recent years [34]; it is also much more computationally expensive. Therefore, it is preferred to improve the parton cascade to yield solutions that are accurate enough without using parton subdivision. We have recently pursued this goal for box calculations [261].

In ZPC, one can take different choices or collision schemes to implement the cascade method [109]. With the closest approach criterion for parton scatterings, the closest approach distance is usually calculated in the two-parton center of mass frame. Two partons may collide when their closest approach distance is smaller than σp/π, and at a given global time all such possible collisions in the future are ordered in a collision list with the ordering time of each collision so that they can be carried out sequentially. The collision list is updated continuously after each collision, and for expansion cases the parton system dynamically freezes out when the collision list is empty. For calculations of a parton system in a box, we terminate the parton cascade at a global time that is large enough so that the parton momentum distribution changes little afterwards. When the closest approach distance is calculated in the two-parton center of mass frame, the collision time of a scattering is a well-defined single value. However, because of the finite σp the two partons have different spatial coordinates in general; thus this collision time in the two-parton center of mass frame becomes two different collision times in the global frame (named here as ct1 and ct2 respectively for the two colliding partons) after the Lorentz transformation. The default collision scheme of ZPC [109] uses (ct1+ct2)/2 as both the collision time and ordering time; this is the case for the AMPT model [13]

Results from the default ZPC scheme [261] at σp=2.6 mb are shown in Fig. 19 (curves with open circles). Panel (a) shows the final parton pT distribution, while panels (b) and (c) show the time evolution of parton 〈pT〉 (scaled by T) and variance of pT (scaled by T2), respectively. The gluon system is initialized in a box with an off-equilibrium initial momentum distribution as shown by the dot-dashed curve in panel (a), where the gluon density is set the same as that for a thermalized gluon system with the Boltzmann distribution at temperature T=0.5 GeV. We see from Fig. 19(a) that the final distribution from the default ZPC scheme deviates considerably from the expected thermal distribution (dotted curve). On the other hand, we find that a new collision scheme, which uses (ct1,ct2) as both the collision time and ordering time, gives a final distribution very close to the thermal distribution [261]. The causality violation usually suppresses collision rates, which is the case for the default ZPC scheme; it is therefore understandable that choosing time (ct1,ct2) instead of (ct1+ct2)/2 enhances the collision rates and thus suppresses the causality violation.

Fig. 19
(Color online)(a) The final pT distribution, (b) time evolution of 〈pT /T〉, and (c) time evolution of 〈pT variance〉/T2 from ZPC from the default collision scheme (open circles) and the new collision scheme (filled squares) in comparison with parton subdivision results at subdivision factor l=106 (dashed curves) for gluons in a box at T=0.5 GeV and σp=2.6 mb.
pic

We use the parton subdivision method to obtain the “exact” time evolutions of 〈pT〉 and pT variance (dashed curves) in Figs. 19(b) and (c). We see that the time evolution of the pT variance from the default scheme deviates significantly from the “exact” parton subdivision result, although the time evolutions of 〈pT〉 are close to each other (mostly due to the conservation of total momentum). In contrast, the time evolution of the pT variance from the new scheme [261] is very close to the parton subdivision result, which at late times agrees with theoretical expectation (diamond). By examining cases of different parton densities and cross sections [261], we find rather surprisingly that the new scheme for ZPC gives very accurate results (i.e., very close to parton subdivision results and/or theoretical values) even at very large opacities, such as the case of T=0.7 GeV and σp=10 mb.

We have used a novel parton subdivision method for the results shown in Fig. 19. In the standard method, one increases the initial parton number per event by factor l while decreasing the cross section by the same factor, which can be schematically represented by the following:

Nl×N, V unchanged, (31)

where N is the initial parton number in an event and V is the initial volume of the parton system. Since the number of possible collisions scales with l2, the subdivision method is very expensive in terms of the computation time, which roughly scales with l2 per subdivision event or l per simulated parton. However, for box calculations where the density function f(x,p,t) is spatially homogeneous, the following new subdivision method can be used:

N unchanged, VV/l, (32)

where we decrease the volume of the box by factor l while keeping the same parton number and momentum distribution in each event. This subdivision method is much more efficient than the standard subdivision method; we therefore use a huge subdivision factor 106 (instead of the usual value of up to a few hundreds).

We emphasize that the differential cross section must not be changed when performing parton subdivision; as a result, the exact transformation for parton subdivision is [261]

f(x,p,t)l×f(x,p,t), dσpdt^dσpdt^/l. (33)

This is especially relevant for forward-angle scattering. For example, when parton subdivision requires the decrease of the forward-angle cross section of Eq.(29), one should not do that by increasing the screening mass μ by a factor of l because that would change the angular distribution of the scatterings in Eq.(28). Instead, one can decrease the αs parameter by a factor of l, which decreases the total scattering cross section while keeping its angular distribution the same.

Transport coefficients such as the shear viscosity η represent important properties of the created matter [262]. Therefore, we have also evaluated the effect of the new collision scheme on the shear viscosity η and its ratio over the entropy density η/s. The Green-Kubo relation [263, 264] has been applied [265-269] to calculate the shear viscosity at or near equilibrium. We thus start with an equilibrium initial condition for shear viscosity calculations according to the Green-Kubo relation [265].

Figure 20 shows our η/s results as functions of the opacity parameter χ, which is defined as [254]

Fig. 20
(Color online) The η/s ratio for different cases of gluon scatterings in a box versus the opacity χ; the solid curve without symbols represents the Navier-Stokes expectation for isotropic scatterings.
pic
χ=σpπ/λ=nσp3π, (34)

where n is the parton density and λ is the mean free path. The case shown in Fig. 19 for gluons in a box at T=0.5 GeV and σp=2.6 mb then corresponds to χ=2.0, and other χ values shown in Fig. 20 are obtained for the following cases: T=0.2 GeV and σp=2.6 mb, T=0.7 GeV and σp=5.2 mb, and T=0.7 GeV and σp=10 mb [261]. For isotropic scatterings of a massless Maxwell-Boltzmann gluon gas in equilibrium (where s=4n and degeneracy factor dg=16), we have the following Navier-Stokes expectation:

(ηs)NS0.4633dg1/3χ2/3=0.1839χ2/3, (35)

which only depends on the opacity χ. We see in Fig. 20 that for isotropic scatterings the subdivision result agrees well with the Navier-Stokes expectation (solid curve). On the other hand, the extracted η and η/s values from the default ZPC scheme are significantly different from the Navier-Stokes expectation or the parton subdivision results at large opacities, although they agree at low opacities as expected. We also see that the results from the new collision scheme are very close to the subdivision results for both forward-angle scatterings and isotropic scatterings, even at a huge opacity χ=41. The new ZPC collision scheme is the first step towards the validation and improvement of the ZPC parton cascade for scatterings in 3-dimensional expansion cases.

3 Other developments

There are other developments of the AMPT model that have not been covered in the previous section. Here we gave a brief overview of some of these works.

The AMPT model has been extended to include deformed nuclei as the projectile and/or target. First, deformed uranium nuclei are implemented [270] to study various observables in U+U collisions at 200A GeV and the effect of nuclear deformation. Later, the AMPT model has been modified to specify the initial proton and neutron spatial distributions in the 96Ru or 96Zr nucleus according to the density functional theory (DFT) calculations [271-273]. The effects of the DFT nuclear density distributions on the backgrounds and possible signals of the chiral magnetic effect (CME) in isobar collisions are then investigated [271]. The extended AMPT model is also used in the study that proposes a novel method to search for the CME in a single heavy ion collision system [272]. Another study [273] uses the model to study multiplicity distributions and elliptic flow in isobar collisions, where the differences between the two isobar systems have the potential to decisively discriminate DFT nuclear distributions from the usual Woods-Saxon density distributions.

The AMPT model has also been extended to include mean field potentials in the hadronic phase in a study of the elliptic flow splitting of particles and antiparticles at the RHIC BES energies [274]. A later study coupled the AMPT model with a parton transport based on the 3-flavor Nambu-Jona-Lasinio model [275] to include the partonic mean field potentials; it shows that a combination of partonic and hadronic mean field potentials can describe the observed splitting of elliptic flows.

The current AMPT model has been known to violate the electric charge conservation because of two reasons [276]. First, the hadron cascade is based on the ART model [84] that has K+ and K- as explicit particles but not K0 or K¯0. As a result, we change K0 to K+ and change 0 to K- prior to the hadron cascade in order to include hadronic interactions of neutral kaons, and after the hadron cascade we assume the isospin symmetry and thus change half of the final K+ into K0 and change half of the final K- to 0. The second reason is that many hadron reactions and some resonance decays in AMPT violate the electric charge conservation. Some reaction channels do not consider electric charges of the initial-state hadrons; instead the isospin-averaged cross section is used and the electric charge of each final state hadron is set randomly [276]. We have developed a version of the AMPT model that has corrected these problems and thus satisfies the electric charge conservation [277]. This charge-conserved version of the AMPT model has been shared with some colleagues for their recent studies of charge-dependent CME signals [278, 279].

Recently we have developed a pure hadron cascade version of the AMPT model (AMPT-HC) [280] to study heavy ion collisions at low energies below a few GeVs. Note that the Eikonal formalism, which is a basis of the HIJING model and thus the initial condition of the standard AMPT model, is expected to break down for nuclear collisions at low enough energies. We thus treat a heavy ion collision as individual nucleon-nucleon collisions in the AMPT-HC model. First, we use the Woods-Saxon nucleon density distribution and the local Thomas-Fermi approximation to initialize the position and momentum of each nucleon in the incoming nuclei. Then primary nucleon-nucleon collisions are treated with the hadron cascade component of AMPT, without going through the Lund string fragmentation, the parton cascade, or quark coalescence. In addition to the usual elastic and inelastic collisions, the hadron cascade in the AMPT-HC model also includes hadron mean field potentials for kaons, baryons and antibaryons. This model has been used to study the Ξ production in low energy Au+Au collisions, which is proposed as a better probe of the nuclear equation of state at high densities than single strangeness (kaon or Λ) productions [280].

4 Summary and outlook

A multi-phase transport model was constructed to provide a self-contained kinetic theory-based description of relativistic nuclear collisions with its four main components: the fluctuating initial condition, partonic interactions, hadronization, and hadronic interactions. Here we review the main developments since the public release of the AMPT source code in 2004 and the 2005 publication that described the details of the model at that time. Several developments have been carried out to improve the initial condition, including the incorporation of finite nuclear thickness relevant for heavy ion collisions below the energy of tens of GeVs, the incorporation of modern parton distribution functions of nuclei for high energy heavy ion collisions, improvement of heavy quark productions, the use of local nuclear scaling of key input parameters for the system size dependence and centrality dependence, and the incorporation of PYTHIA8 and sub-nucleon structure. There are also ongoing efforts to improve the accuracy of the parton cascade without using the parton subdivision method that would alter event-by-event correlations and fluctuations. In addition, the spatial quark coalescence model has been further developed to allow a quark the freedom to form either a meson or a baryon depending on the distance to its coalescing partner(s), which improves baryon and antibaryon productions of the model. Furthermore, deuteron production and annihilation processes have been included in the hadron cascade, an AMPT version that satisfies the electric charge conservation has been developed, and a pure hadron cascade version of the AMPT model is recently developed to study heavy ion collisions at low energies below a few GeVs. For high energy nuclear collisions where the quark-gluon plasma is expected, the string melting version of the AMPT model can now reasonably and simultaneously describe the yield, transverse momentum spectrum and elliptic flow of the bulk matter from small to large collision systems. Consequently, the AMPT model has been applied to the study of various observables in nuclear collisions such as particle yields, particle correlations and anisotropic flows, vorticity and polarization.

Because the transport model approach can address non-equilibrium dynamics, it provides a complementary framework to hydrodynamical models for large systems at high energies, and more importantly it is well suited to study the transition from the dilute limit to the hydrodynamic limit. Therefore, it will be worthwhile to further develop a multi-phase transport as a dynamical model for relativistic nuclear collisions.

There are multiple areas that should be addressed in the future. Regarding the initial condition, at low enough energies the pure hadron cascade version should be applicable while at high enough energies the Eikonal formalism should be valid. It would be desirable to have a unified physics formulation that self-consistently changes from one regime to the other as the colliding energy increases. In addition, for high enough energies and/or large enough collision systems the QGP is expected to be formed, and consequently the string melting version of the AMPT model should be applicable instead the string-dominated default version. The AMPT model should be improved to dynamically determine whether the QGP should be formed in the initial state; it would then self-consistently change from a string-dominated initial condition to a parton-dominated one when the initial energy density is high enough. Another deficiency in the initial condition of the string melting AMPT model is the lack of gluons in the parton phase, and the color-glass-condensate approach would be ideal for including initial gluons once the approach can be generalized to address the quark degrees of freedom such as the nonzero net-baryon number. Regarding the parton phase, the parton cascade should be generalized to perform transport in the presence of an electromagnetic field to enable studies of the electromagnetic field and related observables. Another area of development concerns the study of high net-baryon density physics and the QCD critical point. The AMPT model could be coupled to or improved with effective theories such as the functional renormalization group method or the Nambu-Jona-Lasinio model to treat parton interactions self-consistently including the effective equation of state and effects from the critical point. Regarding the hadronization process, a dynamical parton recombination criterion, e.g., by using the local parton energy density as the recombination criterion instead of starting hadronization at the parton kinetic freezeout, should be developed. Also, additional mechanisms such as independent fragmentation should be included to treat partons that do not find suitable coalescence partners within the local phase space; this would enable the AMPT model to be applicable to studies of high pT physics once the radiative energy loss of high pT partons is considered in the parton phase. Regarding the hadron cascade, it can benefit from the inclusion of more resonances for more realistic thermodynamic properties and chemical equilibration of the hadron matter, and modern models such as the SMASH model could be a good choice as the new hadron cascade component. We expect that the AMPT model in the near future, even if only improved in a few focused areas, will enable us to address some key questions in heavy ion physics and also serve as a more reliable open source transport model for the community for various studies of nuclear collisions.

References
[1] T. D. Lee and G. C. Wick,

Vacuum stability and vacuum excitation in a spin 0 field theory

. Phys. Rev. D 9, 2291-2316 (1974). doi: 10.1103/PhysRevD.9.2291
Baidu ScholarGoogle Scholar
[2] E. V. Shuryak,

Quark-gluon plasma and hadronic production of leptons, photons and psions

. Phys. Lett. B 78, 150 (1978). doi: 10.1016/0370-2693(78)90370-2
Baidu ScholarGoogle Scholar
[3] I. Arsene et al. [BRAHMS],

Quark gluon plasma and color glass condensate at RHIC? The Perspective from the BRAHMS experiment

. Nucl. Phys. A 757, 1-27 (2005). doi: 10.1016/j.nuclphysa.2005.02.130
Baidu ScholarGoogle Scholar
[4] B. B. Back et al. [PHOBOS],

The PHOBOS perspective on discoveries at RHIC

. Nucl. Phys. A 757, 28-101 (2005). doi: 10.1016/j.nuclphysa.2005.03.084
Baidu ScholarGoogle Scholar
[5] J. Adams et al. [STAR],

Experimental and theoretical challenges in the search for the quark gluon plasma: The STAR Collaboration’s critical assessment of the evidence from RHIC collisions

. Nucl. Phys. A 757, 102-183 (2005). doi: 10.1016/j.nuclphysa.2005.03.085
Baidu ScholarGoogle Scholar
[6] K. Adcox et al. [PHENIX],

Formation of dense partonic matter in relativistic nucleus-nucleus collisions at RHIC: Experimental evaluation by the PHENIX collaboration

. Nucl. Phys. A 757, 184-283 (2005). doi: 10.1016/j.nuclphysa.2005.03.086
Baidu ScholarGoogle Scholar
[7] U. Heinz and R. Snellings,

Collective flow and viscosity in relativistic heavy-ion collisions

. Ann. Rev. Nucl. Part. Sci. 63, 123-151 (2013). doi: 10.1146/annurev-nucl-102212-170540
Baidu ScholarGoogle Scholar
[8] W. Busza, K. Rajagopal and W. van der Schee,

Heavy ion collisions: the big picture, and the big questions

. Ann. Rev. Nucl. Part. Sci. 68, 339-376 (2018). doi: 10.1146/annurev-nucl-101917-020852
Baidu ScholarGoogle Scholar
[9] M. Gyulassy and L. McLerran,

New forms of QCD matter discovered at RHIC

. Nucl. Phys. A 750, 30-63 (2005). doi: 10.1016/j.nuclphysa.2004.10.034
Baidu ScholarGoogle Scholar
[10] S. A. Bass et al.,

Microscopic models for ultrarelativistic heavy ion collisions

. Prog. Part. Nucl. Phys. 41, 255-369 (1998). doi: 10.1016/S0146-6410(98)00058-1
Baidu ScholarGoogle Scholar
[11] B. Zhang, C. M. Ko, B. A. Li and Z. W. Lin,

A multiphase transport model for nuclear collisions at RHIC

. Phys. Rev. C 61, 067901 (2000). doi: 10.1103/PhysRevC.61.067901
Baidu ScholarGoogle Scholar
[12] Z. Xu and C. Greiner,

Thermalization of gluons in ultrarelativistic heavy ion collisions by including three-body interactions in a parton cascade

. Phys. Rev. C 71, 064901 (2005). doi: 10.1103/PhysRevC.71.064901
Baidu ScholarGoogle Scholar
[13] Z. W. Lin, C. M. Ko, B. A. Li, B. Zhang and S. Pal,

A Multi-phase transport model for relativistic heavy ion collisions

. Phys. Rev. C 72, 064901 (2005). doi: 10.1103/PhysRevC.72.064901
Baidu ScholarGoogle Scholar
[14] W. Cassing and E. L. Bratkovskaya,

Parton-Hadron-String Dynamics: an off-shell transport approach for relativistic energies

. Nucl. Phys. A 831, 215-242 (2009). doi: 10.1016/j.nuclphysa.2009.09.007
Baidu ScholarGoogle Scholar
[15] P. Huovinen, P. F. Kolb, U. W. Heinz, P. V. Ruuskanen and S. A. Voloshin,

Radial and elliptic flow at RHIC: Further predictions

. Phys. Lett. B 503, 58-64 (2001). doi: 10.1016/S0370-2693(01)00219-2
Baidu ScholarGoogle Scholar
[16] B. Betz, J. Noronha, G. Torrieri, M. Gyulassy, I. Mishustin and D. H. Rischke,

Universality of the diffusion wake from stopped and punch-through jets in heavy-ion collisions

. Phys. Rev. C 79, 034902 (2009). doi: 10.1103/PhysRevC.79.034902
Baidu ScholarGoogle Scholar
[17] B. Schenke, S. Jeon and C. Gale,

Elliptic and triangular flow in event-by-event (3+1)D viscous hydrodynamics

. Phys. Rev. Lett. 106, 042301 (2011). doi: 10.1103/PhysRevLett.106.042301
Baidu ScholarGoogle Scholar
[18] P. Bozek,

Collective flow in p-Pb and d-Pd collisions at TeV energies

. Phys. Rev. C 85, 014911 (2012). doi: 10.1103/PhysRevC.85.014911
Baidu ScholarGoogle Scholar
[19] H. Petersen, J. Steinheimer, G. Burau, M. Bleicher and H. Stöcker,

A fully integrated transport approach to heavy ion reactions with an intermediate hydrodynamic stage

. Phys. Rev. C 78, 044901 (2008). doi: 10.1103/PhysRevC.78.044901
Baidu ScholarGoogle Scholar
[20] K. Werner, I. Karpenko, T. Pierog, M. Bleicher and K. Mikhailov,

Event-by-event simulation of the three-dimensional hydrodynamic evolution from flux tube initial conditions in ultrarelativistic heavy ion collisions

. Phys. Rev. C 82, 044904 (2010). doi: 10.1103/PhysRevC.82.044904
Baidu ScholarGoogle Scholar
[21] H. Song, S. A. Bass, U. Heinz, T. Hirano and C. Shen,

200 A GeV Au+Au collisions serve a nearly perfect quark-gluon liquid

. Phys. Rev. Lett. 106, 192301 (2011). [erratum: Phys. Rev. Lett. 109, 139904 (2012)] doi: 10.1103/PhysRevLett.106.192301
Baidu ScholarGoogle Scholar
[22] Z. W. Lin, S. Pal, C. M. Ko, B. A. Li and B. Zhang,

Charged particle rapidity distributions at relativistic energies

. Phys. Rev. C 64, 011902 (2001). doi: 10.1103/PhysRevC.64.011902
Baidu ScholarGoogle Scholar
[23] X. N. Wang and M. Gyulassy,

HIJING: A Monte Carlo model for multiple jet production in p p, p A and A A collisions

. Phys. Rev. D 44, 3501-3516 (1991). doi: 10.1103/PhysRevD.44.3501
Baidu ScholarGoogle Scholar
[24] M. Gyulassy and X. N. Wang,

HIJING 1.0: A Monte Carlo program for parton and particle production in high-energy hadronic and nuclear collisions

. Comput. Phys. Commun. 83, 307 (1994). doi: 10.1016/0010-4655(94)90057-4
Baidu ScholarGoogle Scholar
[25] T. Sjostrand,

High-energy physics event generation with PYTHIA 5.7 and JETSET 7.4

. Comput. Phys. Commun. 82, 74-90 (1994). doi: 10.1016/0010-4655(94)90132-5
Baidu ScholarGoogle Scholar
[26] Z. W. Lin and C. M. Ko,

Partonic effects on the elliptic flow at RHIC

. Phys. Rev. C 65, 034904 (2002). doi: 10.1103/PhysRevC.65.034904
Baidu ScholarGoogle Scholar
[27] Z. W. Lin,

Evolution of transverse flow and effective temperatures in the parton phase from a multi-phase transport model

. Phys. Rev. C 90, no.1, 014904 (2014). doi: 10.1103/PhysRevC.90.014904
Baidu ScholarGoogle Scholar
[28] D. Molnar and S. A. Voloshin,

Elliptic flow at large transverse momenta from quark coalescence

. Phys. Rev. Lett. 91, 092301 (2003). doi: 10.1103/PhysRevLett.91.092301
Baidu ScholarGoogle Scholar
[29] D. Molnar,

How AMPT generates large elliptic flow with small cross sections

. arXiv:1906.12313 [nucl-th].
Baidu ScholarGoogle Scholar
[30] A. Bzdak and G. L. Ma,

Elliptic and triangular flow in p+Pb and peripheral Pb+Pb collisions from parton scatterings

. Phys. Rev. Lett. 113, no.25, 252301 (2014). doi: 10.1103/PhysRevLett.113.252301
Baidu ScholarGoogle Scholar
[31] G. L. Ma and Z. W. Lin,

Predictions for snn=5.02 TeV Pb+Pb collisions from a multi-phase transport model

. Phys. Rev. C 93, no.5, 054911 (2016). doi: 10.1103/PhysRevC.93.054911
Baidu ScholarGoogle Scholar
[32] Y. He and Z. W. Lin,

Improved quark coalescence for a multi-phase transport model

. Phys. Rev. C 96, no.1, 014910 (2017). doi: 10.1103/PhysRevC.96.014910
Baidu ScholarGoogle Scholar
[33] L. Y. Zhang, J. H. Chen, Z. W. Lin, Y. G. Ma and S. Zhang,

Two-particle angular correlations in pp and p-Pb collisions at energies available at the CERN Large Hadron Collider from a multiphase transport model

. Phys. Rev. C 98, no.3, 034912 (2018). doi: 10.1103/PhysRevC.98.034912
Baidu ScholarGoogle Scholar
[34] B. Alver and G. Roland,

Collision geometry fluctuations and triangular flow in heavy-ion collisions

. Phys. Rev. C 81, 054905 (2010). [erratum: Phys. Rev. C 82, 039903 (2010)] doi: 10.1103/PhysRevC.82.039903
Baidu ScholarGoogle Scholar
[35] Y. Jiang, Z. W. Lin and J. Liao,

Rotating quark-gluon plasma in relativistic heavy ion collisions

. Phys. Rev. C 94, no.4, 044910 (2016). [erratum: Phys. Rev. C 95, no.4, 049904 (2017)] doi: 10.1103/PhysRevC.94.044910
Baidu ScholarGoogle Scholar
[36] H. Li, L. G. Pang, Q. Wang and X. L. Xia,

Global Λ polarization in heavy-ion collisions from a transport model

. Phys. Rev. C 96, no.5, 054908 (2017). doi: 10.1103/PhysRevC.96.054908
Baidu ScholarGoogle Scholar
[37] S. Lan, Z. W. Lin, S. Shi and X. Sun,

Effects of finite coverage on global polarization observables in heavy ion collisions

. Phys. Lett. B 780, 319-324 (2018). doi: 10.1016/j.physletb.2018.02.076
Baidu ScholarGoogle Scholar
[38] B. I. Abelev et al. [STAR],

Systematic measurements of identified particle spectra in p p, d+ Au and Au+Au collisions from STAR

. Phys. Rev. C 79, 034909 (2009). doi: 10.1103/PhysRevC.79.034909
Baidu ScholarGoogle Scholar
[39] J. Y. Ollitrault,

Anisotropy as a signature of transverse collective flow

. Phys. Rev. D 46, 229-245 (1992). doi: 10.1103/PhysRevD.46.229
Baidu ScholarGoogle Scholar
[40] C. Gale, S. Jeon and B. Schenke,

Hydrodynamic modeling of heavy-ion collisions

. Int. J. Mod. Phys. A 28, 1340011 (2013). doi: 10.1142/S0217751X13400113
Baidu ScholarGoogle Scholar
[41] H. Li, L. He, Z. W. Lin, D. Molnar, F. Wang and W. Xie,

Origin of the mass splitting of elliptic anisotropy in a multiphase transport model

. Phys. Rev. C 93, no.5, 051901 (2016). doi: 10.1103/PhysRevC.93.051901
Baidu ScholarGoogle Scholar
[42] H. Li, L. He, Z. W. Lin, D. Molnar, F. Wang and W. Xie,

Origin of the mass splitting of azimuthal anisotropies in a multiphase transport model

. Phys. Rev. C 96, no.1, 014901 (2017). doi: 10.1103/PhysRevC.96.014901
Baidu ScholarGoogle Scholar
[43] D. Molnar and M. Gyulassy,

Saturation of elliptic flow and the transport opacity of the gluon plasma at RHIC

. Nucl. Phys. A 697, 495-520 (2002). [erratum: Nucl. Phys. A 703, 893-894 (2002)]. doi: 10.1016/S0375-9474(01)01224-6
Baidu ScholarGoogle Scholar
[44] V. Khachatryan et al. [CMS],

Observation of long-range near-side angular correlations in proton-proton collisions at the LHC

. JHEP 09, 091 (2010). doi: 10.1007/JHEP09(2010)091
Baidu ScholarGoogle Scholar
[45] A. Adare et al. [PHENIX],

Measurement of long-range angular correlation and quadrupole anisotropy of pions and (anti)protons in central d+Au collisions at sNN=200 GeV

. Phys. Rev. Lett. 114, no.19, 192301 (2015). doi: 10.1103/PhysRevLett.114.192301
Baidu ScholarGoogle Scholar
[46] C. Aidala et al. [PHENIX],

Creation of quark–gluon plasma droplets with three distinct geometries

. Nature Phys. 15, no.3, 214-220 (2019). doi: 10.1038/s41567-018-0360-0
Baidu ScholarGoogle Scholar
[47] P. Bozek,

Elliptic flow in proton-proton collisions at sqrt(S) = 7 TeV

. Eur. Phys. J. C 71, 1530 (2011). doi: 10.1140/epjc/s10052-010-1530-0
Baidu ScholarGoogle Scholar
[48] P. Bozek and W. Broniowski,

Correlations from hydrodynamic flow in p-Pb collisions

. Phys. Lett. B 718, 1557-1561 (2013). doi: 10.1016/j.physletb.2012.12.051
Baidu ScholarGoogle Scholar
[49] L. He, T. Edmonds, Z. W. Lin, F. Liu, D. Molnar and F. Wang,

Anisotropic parton escape is the dominant source of azimuthal anisotropy in transport models

. Phys. Lett. B 753, 506-510 (2016). doi: 10.1016/j.physletb.2015.12.051
Baidu ScholarGoogle Scholar
[50] Z. W. Lin, L. He, T. Edmonds, F. Liu, D. Molnar and F. Wang,

Elliptic anisotropy v2 may be dominated by particle escape instead of hydrodynamic flow

. Nucl. Phys. A 956, 316-319 (2016). doi: 10.1016/j.nuclphysa.2016.01.017
Baidu ScholarGoogle Scholar
[51] A. Kurkela, U. A. Wiedemann and B. Wu,

Nearly isentropic flow at sizeable η/s

. Phys. Lett. B 783, 274-279 (2018). doi: 10.1016/j.physletb.2018.06.064
Baidu ScholarGoogle Scholar
[52] U. W. Heinz and J. S. Moreland,

Hydrodynamic flow in small systems or: “How the heck is it possible that a system emitting only a dozen particles can be described by fluid dynamics?”

. J. Phys. Conf. Ser. 1271, no.1, 012018 (2019). doi: 10.1088/1742-6596/1271/1/012018
Baidu ScholarGoogle Scholar
[53] B. Schenke,

The smallest fluid on earth

. arXiv:2102.11189 [nucl-th].
Baidu ScholarGoogle Scholar
[54] K. Dusling, M. Mace and R. Venugopalan,

Multiparticle collectivity from initial state correlations in high energy proton-nucleus collisions

. Phys. Rev. Lett. 120, no.4, 042002 (2018). doi: 10.1103/PhysRevLett.120.042002
Baidu ScholarGoogle Scholar
[55] M. Mace, V. V. Skokov, P. Tribedy and R. Venugopalan,

Hierarchy of azimuthal anisotropy harmonics in collisions of small systems from the color glass condensate

. Phys. Rev. Lett. 121, no.5, 052301 (2018). [erratum: Phys. Rev. Lett. 123, no.3, 039901 (2019)] doi: 10.1103/PhysRevLett.121.052301
Baidu ScholarGoogle Scholar
[56] R. D. Weller and P. Romatschke,

One fluid to rule them all: viscous hydrodynamic description of event-by-event central p+p, p+Pb and Pb+Pb collisions at s=5.02 TeV

. Phys. Lett. B 774, 351-356 (2017). doi: 10.1016/j.physletb.2017.09.077
Baidu ScholarGoogle Scholar
[57] A. Kurkela, U. A. Wiedemann and B. Wu,

Flow in AA and pA as an interplay of fluid-like and non-fluid like excitations

. Eur. Phys. J. C 79, no.11, 965 (2019). doi: 10.1140/epjc/s10052-019-7428-6
Baidu ScholarGoogle Scholar
[58] R. A. Lacey [STAR],

Long-range collectivity in small collision-systems with two- and four-particle correlations @ STAR

. Nucl. Phys. A 1005, 122041 (2021). doi: 10.1016/j.nuclphysa.2020.122041
Baidu ScholarGoogle Scholar
[59] U. A. Acharya et al. [PHENIX],

Kinematic dependence of azimuthal anisotropies in p + Au, d + Au, 3He + Au at sNN = 200 GeV

. arXiv:2107.06634 [hep-ex].
Baidu ScholarGoogle Scholar
[60]

The Open Standard Codes and Routines (OSCAR)

project at https://karman.physics.purdue.edu/OSCAR-old/
Baidu ScholarGoogle Scholar
[61]

Source codes of various AMPT

versions are available at http://myweb.ecu.edu/linz/ampt/
Baidu ScholarGoogle Scholar
[62] B. I. Abelev et al. [STAR],

Observation of an antimatter hypernucleus

. Science 328, 58-62 (2010). doi: 10.1126/science.1183980
Baidu ScholarGoogle Scholar
[63] N. Sharma [ALICE],

Production of nuclei and antinuclei in pp and Pb-Pb collisions with ALICE at the LHC

. J. Phys. G 38, 124189 (2011). doi: 10.1088/0954-3899/38/12/124189
Baidu ScholarGoogle Scholar
[64] K. J. Sun, L. W. Chen, C. M. Ko and Z. Xu,

Probing QCD critical fluctuations from light nuclei production in relativistic heavy-ion collisions

. Phys. Lett. B 774, 103-107 (2017). doi: 10.1016/j.physletb.2017.09.056
Baidu ScholarGoogle Scholar
[65] E. Shuryak and J. M. Torres-Rincon,

Baryon clustering at the critical line and near the hypothetical critical point in heavy-ion collisions

. Phys. Rev. C 100, no.2, 024903 (2019). doi: 10.1103/PhysRevC.100.024903
Baidu ScholarGoogle Scholar
[66] A. Bzdak, S. Esumi, V. Koch, J. Liao, M. Stephanov and N. Xu,

Mapping the phases of quantum chromodynamics with beam energy scan

. Phys. Rept. 853, 1-87 (2020). doi: 10.1016/j.physrep.2020.01.005
Baidu ScholarGoogle Scholar
[67] X. Luo, S. Shi, N. Xu and Y. Zhang,

A study of the properties of the QCD phase diagram in high-energy nuclear collisions

. Particles 3, no.2, 278-307 (2020). doi: 10.3390/particles3020022
Baidu ScholarGoogle Scholar
[68] J. Cleymans and H. Satz,

Thermal hadron production in high-energy heavy ion collisions

. Z. Phys. C 57, 135-148 (1993). doi: 10.1007/BF01555746
Baidu ScholarGoogle Scholar
[69] P. Braun-Munzinger, V. Koch, T. Schäfer and J. Stachel,

Properties of hot and dense matter from relativistic heavy ion collisions

. Phys. Rept. 621, 76-126 (2016). doi: 10.1016/j.physrep.2015.12.003
Baidu ScholarGoogle Scholar
[70] H. Sato and K. Yazaki,

On the coalescence model for high-energy nuclear reactions

. Phys. Lett. B 98, 153-157 (1981). doi: 10.1016/0370-2693(81)90976-X
Baidu ScholarGoogle Scholar
[71] L. P. Csernai and J. I. Kapusta,

Entropy and cluster production in nuclear collisions

. Phys. Rept. 131, 223-318 (1986). doi: 10.1016/0370-1573(86)90031-1
Baidu ScholarGoogle Scholar
[72] H. H. Gutbrod, A. Sandoval, P. J. Johansen, A. M. Poskanzer, J. Gosset, W. G. Meyer, G. D. Westfall and R. Stock,

Final state interactions in the production of hydrogen and helium isotopes by relativistic heavy ions on uranium

. Phys. Rev. Lett. 37, 667-670 (1976). doi: 10.1103/PhysRevLett.37.667
Baidu ScholarGoogle Scholar
[73] L. Zhu, C. M. Ko and X. Yin,

Light (anti-)nuclei production and flow in relativistic heavy-ion collisions

. Phys. Rev. C 92, no.6, 064911 (2015). doi: 10.1103/PhysRevC.92.064911
Baidu ScholarGoogle Scholar
[74] K. J. Sun and L. W. Chen,

Analytical coalescence formula for particle production in relativistic heavy-ion collisions

. Phys. Rev. C 95, no.4, 044905 (2017). doi: 10.1103/PhysRevC.95.044905
Baidu ScholarGoogle Scholar
[75] P. Danielewicz and G. F. Bertsch,

Production of deuterons and pions in a transport model of energetic heavy ion reactions

. Nucl. Phys. A 533, 712-748 (1991). doi: 10.1016/0375-9474(91)90541-D
Baidu ScholarGoogle Scholar
[76] Y. Oh and C. M. Ko,

Elliptic flow of deuterons in relativistic heavy-ion collisions

. Phys. Rev. C 76, 054910 (2007). doi: 10.1103/PhysRevC.76.054910
Baidu ScholarGoogle Scholar
[77] Y. Oh, Z. W. Lin and C. M. Ko,

Deuteron production and elliptic flow in relativistic heavy ion collisions

. Phys. Rev. C 80, 064902 (2009). doi: 10.1103/PhysRevC.80.064902
Baidu ScholarGoogle Scholar
[78] R. M. Heinz, O. E. Overseth, D. E. Pellett and M. L. Perl,

Differential Cross Sections for p+p – > d+pi+ from 1 to 3 BeV

. Phys. Rev. 167, 1232-1239 (1968). doi: 10.1103/PhysRev.167.1232
Baidu ScholarGoogle Scholar
[79] H. L. Anderson, D. A. Larson, L. C. Myrianthopoulos, L. Dubal, C. K. Hargrove, E. P. Hincks, R. J. Mckee, H. Mes, D. Kessler and A. C. Thompson,

Measurements of the differential cross-section of the reaction p p → d pi+ from 3.0 to 5.0 gev/c

. Phys. Rev. D 9, 580-596 (1974). doi: 10.1103/PhysRevD.9.580
Baidu ScholarGoogle Scholar
[80] F. Shimizu, Y. Kubota, H. Koiso, F. Sai, S. Sakamoto and S. S. Yamamoto,

Measurement of the p p cross-sections in the momentum range 0.9-2.0 GeV/c

. Nucl. Phys. A 386, 571-588 (1982). doi: 10.1016/0375-9474(82)90037-9
Baidu ScholarGoogle Scholar
[81] J. Boswell, R. Altemus, R. Minehart, L. Orphanos, H. J. Ziock and E. A. Wadlinger,

Differential cross-section for the pi+ d → p p reaction from 80-MeV to 417-MeV

. Phys. Rev. C 25, 2540-2549 (1982). doi: 10.1103/PhysRevC.25.2540
Baidu ScholarGoogle Scholar
[82] M. Y. Borkovsky et al.,

Measurement of the differential cross-sections for the pi+ d → p p reaction at pion energies of 280-MeV, 300-MeV, 330-MeV, 357-MeV, 390-MeV, 420-MeV and 450-MeV

. J. Phys. G 11, 69-83 (1985). doi: 10.1088/0305-4616/11/1/012
Baidu ScholarGoogle Scholar
[83] S. I. Gogolev et al.,

Total cross-section of the reaction pi+ d → p p at pion energies 26-MeV to 40-MeV

. Phys. Lett. B 300, 24-28 (1993). doi: 10.1016/0370-2693(93)90742-Z
Baidu ScholarGoogle Scholar
[84] B. A. Li and C. M. Ko,

Formation of superdense hadronic matter in high-energy heavy ion collisions

. Phys. Rev. C 52, 2037-2063 (1995). doi: 10.1103/PhysRevC.52.2037
Baidu ScholarGoogle Scholar
[85] B. A. Li and C. M. Ko,

Excitation functions in central Au + Au collisions from SIS / GSI to AGS Brookhaven

. Nucl. Phys. A 601, 457-472 (1996). doi: 10.1016/0375-9474(96)00037-1
Baidu ScholarGoogle Scholar
[86] B. Li, A. T. Sustich, B. Zhang and C. M. Ko,

Studies of superdense hadronic matter in a relativistic transport model

. Int. J. Mod. Phys. E 10, 267-352 (2001). doi: 10.1142/S0218301301000575
Baidu ScholarGoogle Scholar
[87] R. A. Arndt, I. I. Strakovsky, R. L. Workman and D. V. Bugg,

Analysis of the reaction pi+ d → p p to 500-MeV

. Phys. Rev. C 48, 1926-1938 (1993). doi: 10.1103/PhysRevC.49.1229
Baidu ScholarGoogle Scholar
[88] M. Gyulassy, K. Frankel and E. a. Remler,

Deuteron formation in nuclear collisions

. Nucl. Phys. A 402, 596-611 (1983). doi: 10.1016/0375-9474(83)90222-1
Baidu ScholarGoogle Scholar
[89] D. Oliinychenko, L. G. Pang, H. Elfner and V. Koch,

Microscopic study of deuteron production in PbPb collisions at s=2.76TeV via hydrodynamics and a hadronic afterburner

. Phys. Rev. C 99, no.4, 044907 (2019). doi: 10.1103/PhysRevC.99.044907
Baidu ScholarGoogle Scholar
[90] K. J. Sun, C. M. Ko and Z. W. Lin,

Light nuclei production in a multiphase transport model for relativistic heavy ion collisions

. Phys. Rev. C 103, no.6, 064909 (2021). doi: 10.1103/PhysRevC.103.064909
Baidu ScholarGoogle Scholar
[91] K. J. Sun, C. M. Ko and B. Dönigus,

Suppression of light nuclei production in collisions of small systems at the Large Hadron Collider

. Phys. Lett. B 792, 132-137 (2019). doi: 10.1016/j.physletb.2019.03.033
Baidu ScholarGoogle Scholar
[92] B. Andersson, G. Gustafson and B. Soderberg,

A general model for jet fragmentation

. Z. Phys. C 20, 317 (1983). doi: 10.1007/BF01407824
Baidu ScholarGoogle Scholar
[93] B. Andersson, G. Gustafson, G. Ingelman and T. Sjostrand,

Parton fragmentation and string dynamics

. Phys. Rept. 97, 31-145 (1983). doi: 10.1016/0370-1573(83)90080-7
Baidu ScholarGoogle Scholar
[94] Z. W. Lin, C. M. Ko and S. Pal,

Partonic effects on pion interferometry at RHIC

. Phys. Rev. Lett. 89, 152301 (2002). doi: 10.1103/PhysRevLett.89.152301
Baidu ScholarGoogle Scholar
[95] J. Xu and C. M. Ko,

Pb-Pb collisions at sNN=2.76 TeV in a multiphase transport model

. Phys. Rev. C 83, 034904 (2011). doi: 10.1103/PhysRevC.83.034904
Baidu ScholarGoogle Scholar
[96] J. S. Schwinger,

Gauge invariance and mass. II

. Phys. Rev. 128, 2425-2429 (1962). doi: 10.1103/PhysRev.128.2425
Baidu ScholarGoogle Scholar
[97] A. Adare et al. [PHENIX],

Neutral pion production with respect to centrality and reaction plane in Au+Au collisions at sNN=200 GeV

. Phys. Rev. C 87, no.3, 034911 (2013). doi: 10.1103/PhysRevC.87.034911
Baidu ScholarGoogle Scholar
[98] B. Abelev et al. [ALICE],

Centrality determination of Pb-Pb collisions at sNN = 2.76 TeV with ALICE

. Phys. Rev. C 88, no.4, 044909 (2013). doi: 10.1103/PhysRevC.88.044909
Baidu ScholarGoogle Scholar
[99] S. S. Adler et al. [PHENIX],

Identified charged particle spectra and yields in Au+Au collisions at S(NN)**1/2 = 200-GeV

. Phys. Rev. C 69, 034909 (2004). doi: 10.1103/PhysRevC.69.034909
Baidu ScholarGoogle Scholar
[100] I. G. Bearden et al. [BRAHMS],

Charged meson rapidity distributions in central Au+Au collisions at s(NN)**(1/2) = 200-GeV

. Phys. Rev. Lett. 94, 162301 (2005). doi: 10.1103/PhysRevLett.94.162301
Baidu ScholarGoogle Scholar
[101] B. Abelev et al. [ALICE],

Centrality dependence of π, K, p production in Pb-Pb collisions at sNN = 2.76 TeV

. Phys. Rev. C 88, 044910 (2013). doi: 10.1103/PhysRevC.88.044910
Baidu ScholarGoogle Scholar
[102] Y. Gu [PHENIX],

PHENIX measurements of higher-order flow harmonics for identified charged hadrons in Au+Au collisions at sNN=39−200 GeV

. Nucl. Phys. A 904-905, 353c-356c (2013). doi: 10.1016/j.nuclphysa.2013.02.022
Baidu ScholarGoogle Scholar
[103] G. Aad et al. [ATLAS],

Measurement of the azimuthal anisotropy for charged particle production in sNN=2.76 TeV lead-lead collisions with the ATLAS detector

. Phys. Rev. C 86, 014907 (2012). doi: 10.1103/PhysRevC.86.014907
Baidu ScholarGoogle Scholar
[104] S. Chatrchyan et al. [CMS],

Measurement of the elliptic anisotropy of charged particles produced in PbPb collisions at sNN=2.76 TeV

. Phys. Rev. C 87, no.1, 014902 (2013). doi: 10.1103/PhysRevC.87.014902
Baidu ScholarGoogle Scholar
[105] G. Aad et al. [ATLAS],

Measurement of the centrality and pseudorapidity dependence of the integrated elliptic flow in lead-lead collisions at sNN=2.76 TeV with the ATLAS detector

. Eur. Phys. J. C 74, no.8, 2982 (2014). doi: 10.1140/epjc/s10052-014-2982-4
Baidu ScholarGoogle Scholar
[106] B. B. Back et al. [PHOBOS],

Centrality and pseudorapidity dependence of elliptic flow for charged hadrons in Au+Au collisions at s(NN)**(1/2) = 200-GeV

. Phys. Rev. C 72, 051901 (2005). doi: 10.1103/PhysRevC.72.051901
Baidu ScholarGoogle Scholar
[107] V. Khachatryan et al. [CMS],

Evidence for transverse momentum and pseudorapidity dependent event plane fluctuations in PbPb and pPb collisions

. Phys. Rev. C 92, no.3, 034911 (2015). doi: 10.1103/PhysRevC.92.034911
Baidu ScholarGoogle Scholar
[108] L. G. Pang, H. Petersen, G. Y. Qin, V. Roy and X. N. Wang,

Decorrelation of anisotropic flow along the longitudinal direction

. Eur. Phys. J. A 52, no.4, 97 (2016). doi: 10.1140/epja/i2016-16097-x
Baidu ScholarGoogle Scholar
[109] B. Zhang,

ZPC 1.0.1: A Parton cascade for ultrarelativistic heavy ion collisions

. Comput. Phys. Commun. 109, 193-206 (1998). doi: 10.1016/S0010-4655(98)00010-1
Baidu ScholarGoogle Scholar
[110] V. Minissale, F. Scardina and V. Greco,

Hadrons from coalescence plus fragmentation in AA collisions at energies available at the BNL Relativistic Heavy Ion Collider to the CERN Large Hadron Collider

. Phys. Rev. C 92, no.5, 054904 (2015). doi: 10.1103/PhysRevC.92.054904
Baidu ScholarGoogle Scholar
[111] K. C. Han, R. J. Fries and C. M. Ko,

Jet fragmentation via recombination of parton showers

. Phys. Rev. C 93, no.4, 045207 (2016). doi: 10.1103/PhysRevC.93.045207
Baidu ScholarGoogle Scholar
[112] Y. He and Z. W. Lin,

Baryon spectra and antiparticle-to-particle ratios from the improved AMPT model

. EPJ Web Conf. 171, 14004 (2018). doi: 10.1051/epjconf/201817114004
Baidu ScholarGoogle Scholar
[113] L. Y. Zhang, J. H. Chen, Z. W. Lin, Y. G. Ma and S. Zhang,

Two-particle angular correlations in heavy ion collisions from a multiphase transport model

. Phys. Rev. C 99, no.5, 054904 (2019). doi: 10.1103/PhysRevC.99.054904
Baidu ScholarGoogle Scholar
[114] C. Suire [STAR],

Omega- and anti-Omega+ production in Au+Au collisions at s(NN)**(1/2) = 130-GeV and 200-GeV

. Nucl. Phys. A 715, 470-4735 (2003). doi: 10.1016/S0375-9474(02)01451-3
Baidu ScholarGoogle Scholar
[115] J. Adams et al. [STAR],

Scaling properties of hyperon production in Au+Au Collisions at s**(1/2) = 200-GeV

. Phys. Rev. Lett. 98, 062301 (2007). doi: 10.1103/PhysRevLett.98.062301
Baidu ScholarGoogle Scholar
[116] S. Schuchmann, Modification of Ks0 and Λ(Λ¯) transverse momentum spectra in Pb-Pb collisions at sNN = 2.76 TeV with ALICE. doi: 10.1007/978-3-319-43458-2
[117] B. B. Abelev et al. [ALICE],

Multi-strange baryon production at mid-rapidity in Pb-Pb collisions at sNN = 2.76 TeV

. Phys. Lett. B 728, 216-227 (2014). [erratum: Phys. Lett. B 734, 409-410 (2014)]. doi: 10.1016/j.physletb.2014.05.052
Baidu ScholarGoogle Scholar
[118] J. Zimanyi, T. S. Biro, T. Csorgo and P. Levai,

Quark liberation and coalescence at CERN SPS

. Phys. Lett. B 472, 243-246 (2000). doi: 10.1016/S0370-2693(99)01461-6
Baidu ScholarGoogle Scholar
[119] C. Zhang, L. Zheng, F. Liu, S. Shi and Z. W. Lin,

Update of a multiphase transport model with modern parton distribution functions and nuclear shadowing

. Phys. Rev. C 99, no.6, 064906 (2019). doi: 10.1103/PhysRevC.99.064906
Baidu ScholarGoogle Scholar
[120] C. Nonaka, B. Muller, S. A. Bass and M. Asakawa,

Possible resolutions of the D-paradox

. Phys. Rev. C 71, 051901 (2005). doi: 10.1103/PhysRevC.71.051901
Baidu ScholarGoogle Scholar
[121] V. Greco,

Phase-space coalescence for heavy and light quarks at RHIC

. Eur. Phys. J. ST 155, 45-59 (2008). doi: 10.1140/epjst/e2008-00588-y
Baidu ScholarGoogle Scholar
[122] L. Zheng, C. Zhang, S. S. Shi and Z. W. Lin,

Improvement of heavy flavor production in a multiphase transport model updated with modern nuclear parton distribution functions

. Phys. Rev. C 101, no.3, 034905 (2020). doi: 10.1103/PhysRevC.101.034905
Baidu ScholarGoogle Scholar
[123] T. Shao, J. Chen, C. M. Ko and Z. W. Lin,

Enhanced production of strange baryons in high-energy nuclear collisions from a multiphase transport model

. Phys. Rev. C 102, no.1, 014906 (2020). doi: 10.1103/PhysRevC.102.014906
Baidu ScholarGoogle Scholar
[124] Z. W. Lin,

Extension of the Bjorken energy density formula of the initial state for relativistic heavy ion collisions

. Phys. Rev. C 98, no.3, 034908 (2018). doi: 10.1103/PhysRevC.98.034908
Baidu ScholarGoogle Scholar
[125] T. Mendenhall and Z. W. Lin,

Calculating the initial energy density in heavy ion collisions by including the finite nuclear thickness

. Phys. Rev. C 103, no.2, 024907 (2021). doi: 10.1103/PhysRevC.103.024907
Baidu ScholarGoogle Scholar
[126] M. Okai, K. Kawaguchi, Y. Tachibana and T. Hirano,

New approach to initializing hydrodynamic fields and mini-jet propagation in quark-gluon fluids

. Phys. Rev. C 95, no.5, 054914 (2017). doi: 10.1103/PhysRevC.95.054914
Baidu ScholarGoogle Scholar
[127] C. Shen, G. Denicol, C. Gale, S. Jeon, A. Monnai and B. Schenke,

A hybrid approach to relativistic heavy-ion collisions at the RHIC BES energies

. Nucl. Phys. A 967, 796-799 (2017). doi: 10.1016/j.nuclphysa.2017.06.008
Baidu ScholarGoogle Scholar
[128] J. D. Bjorken,

Highly relativistic nucleus-nucleus collisions: the central rapidity region

. Phys. Rev. D 27, 140-151 (1983). doi: 10.1103/PhysRevD.27.140
Baidu ScholarGoogle Scholar
[129]

A web interface that calculates the average initial energy density at mid-pseudorapidity of central AA collisions with the semi-analytical method of Ref. [125]

is available at http://myweb.ecu.edu/linz/densities/
Baidu ScholarGoogle Scholar
[130] E. Eichten, I. Hinchliffe, K. D. Lane and C. Quigg,

Super collider physics

. Rev. Mod. Phys. 56, 579-707 (1984). doi: 10.1103/RevModPhys.56.579
Baidu ScholarGoogle Scholar
[131] T. K. Gaisser and F. Halzen,

Soft hard scattering in the TeV range

. Phys. Rev. Lett. 54, 1754 (1985). doi: 10.1103/PhysRevLett.54.1754
Baidu ScholarGoogle Scholar
[132] G. Pancheri and Y. N. Srivastava,

Low pT jets and the rise with energy of the inelastic cross-section

. Phys. Lett. B 182, 199-207 (1986). doi: 10.1016/0370-2693(86)91577-7
Baidu ScholarGoogle Scholar
[133] S. Forte and G. Watt,

Progress in the determination of the partonic structure of the proton

. Ann. Rev. Nucl. Part. Sci. 63, 291-328 (2013). doi: 10.1146/annurev-nucl-102212-170607
Baidu ScholarGoogle Scholar
[134] A. De Roeck and R. S. Thorne,

Structure functions

. Prog. Part. Nucl. Phys. 66, 727-781 (2011). doi: 10.1016/j.ppnp.2011.06.001
Baidu ScholarGoogle Scholar
[135] K. Kovařík, P. M. Nadolsky and D. E. Soper,

Hadronic structure in high-energy collisions

. Rev. Mod. Phys. 92, no.4, 045003 (2020). doi: 10.1103/RevModPhys.92.045003
Baidu ScholarGoogle Scholar
[136] H. L. Lai, J. Huston, S. Mrenna, P. Nadolsky, D. Stump, W. K. Tung and C. P. Yuan,

Parton distributions for event generators

. JHEP 04, 035 (2010). doi: 10.1007/JHEP04(2010)035
Baidu ScholarGoogle Scholar
[137] J. C. Collins and X. M. Zu,

Parton distribution functions suitable for Monte Carlo event generators

. JHEP 06, 018 (2002). doi: 10.1088/1126-6708/2002/06/018
Baidu ScholarGoogle Scholar
[138] T. Kasemets and T. Sjostrand,

A Comparison of new MC-adapted Parton Densities

. Eur. Phys. J. C 69, 19-29 (2010). doi: 10.1140/epjc/s10052-010-1391-6
Baidu ScholarGoogle Scholar
[139] D. W. Duke and J. F. Owens,

Q**2 dependent parametrizations of parton distribution functions

. Phys. Rev. D 30, 49-54 (1984). doi: 10.1103/PhysRevD.30.49
Baidu ScholarGoogle Scholar
[140] Z. W. Lin,

Current status and further improvements of a multi-phase transport (AMPT) model

. Indian J. Phys. 85, 837-841 (2011). doi: 10.1007/s12648-011-0086-7
Baidu ScholarGoogle Scholar
[141] J. Pumplin, D. R. Stump, J. Huston, H. L. Lai, P. M. Nadolsky and W. K. Tung,

New generation of parton distributions with uncertainties from global QCD analysis

. JHEP 07, 012 (2002). doi: 10.1088/1126-6708/2002/07/012
Baidu ScholarGoogle Scholar
[142] W. T. Deng, X. N. Wang and R. Xu,

Hadron production in p+p, p+Pb, and Pb+Pb collisions with the HIJING 2.0 model at energies available at the CERN Large Hadron Collider

. Phys. Rev. C 83, 014915 (2011). doi: 10.1103/PhysRevC.83.014915
Baidu ScholarGoogle Scholar
[143] M. Gluck, E. Reya and A. Vogt,

Dynamical parton distributions of the proton and small x physics

. Z. Phys. C 67, 433-448 (1995). doi: 10.1007/BF01624586
Baidu ScholarGoogle Scholar
[144] K. J. Eskola, P. Paakkinen, H. Paukkunen and C. A. Salgado,

EPPS16: Nuclear parton distributions with LHC data

. Eur. Phys. J. C 77, no.3, 163 (2017). doi: 10.1140/epjc/s10052-017-4725-9
Baidu ScholarGoogle Scholar
[145] D. de Florian, R. Sassot, P. Zurita and M. Stratmann,

Global analysis of nuclear parton distributions

. Phys. Rev. D 85, 074028 (2012). doi: 10.1103/PhysRevD.85.074028
Baidu ScholarGoogle Scholar
[146] K. Kovarik et al.,

nCTEQ15 - Global analysis of nuclear parton distributions with uncertainties in the CTEQ framework

. Phys. Rev. D 93, no.8, 085037 (2016). doi: 10.1103/PhysRevD.93.085037
Baidu ScholarGoogle Scholar
[147] M. Hirai, S. Kumano and T. H. Nagai,

Determination of nuclear parton distribution functions and their uncertainties in next-to-leading order

. Phys. Rev. C 76, 065207 (2007). doi: 10.1103/PhysRevC.76.065207
Baidu ScholarGoogle Scholar
[148] R. Abdul Khalek et al. [NNPDF],

Nuclear parton distributions from lepton-nucleus scattering and the impact of an electron-ion collider

. Eur. Phys. J. C 79, no.6, 471 (2019). doi: 10.1140/epjc/s10052-019-6983-1
Baidu ScholarGoogle Scholar
[149] K. J. Eskola,

Shadowing effects on quark and gluon production in ultrarelativistic heavy ion collisions

. Z. Phys. C 51, 633-642 (1991). doi: 10.1007/BF01565590
Baidu ScholarGoogle Scholar
[150] V. Emel’yanov, A. Khodinov, S. R. Klein and R. Vogt,

The Effect of shadowing on initial conditions, transverse energy and hard probes in ultrarelativistic heavy ion collisions

. Phys. Rev. C 61, 044904 (2000). doi: 10.1103/PhysRevC.61.044904
Baidu ScholarGoogle Scholar
[151] S. R. Klein and R. Vogt,

Inhomogeneous shadowing effects on J/psi production in dA collisions

. Phys. Rev. Lett. 91, 142301 (2003). doi: 10.1103/PhysRevLett.91.142301
Baidu ScholarGoogle Scholar
[152] L. Frankfurt, V. Guzey and M. Strikman,

Leading twist nuclear shadowing phenomena in hard processes with nuclei

. Phys. Rept. 512, 255-393 (2012). doi: 10.1016/j.physrep.2011.12.002
Baidu ScholarGoogle Scholar
[153] J. L. Nagle, A. D. Frawley, L. A. L. Levy and M. G. Wysocki,

Theoretical modeling of J/psi yield modifications in proton (deuteron) - nucleus collisions at high energy

. Phys. Rev. C 84, 044911 (2011). doi: 10.1103/PhysRevC.84.044911
Baidu ScholarGoogle Scholar
[154] B. Wu,

Factorization and transverse phase-space parton distributions

. arXiv:2102.12916 [hep-ph].
Baidu ScholarGoogle Scholar
[155] I. Helenius, K. J. Eskola, H. Honkanen and C. A. Salgado,

Impact-parameter dependent nuclear parton distribution functions: EPS09s and EKS98s and their applications in nuclear hard processes

. JHEP 07, 073 (2012). doi: 10.1007/JHEP07(2012)073
Baidu ScholarGoogle Scholar
[156] K. J. Eskola, V. J. Kolhinen and P. V. Ruuskanen,

Scale evolution of nuclear parton distributions

. Nucl. Phys. B 535, 351-371 (1998). doi: 10.1016/S0550-3213(98)00589-6
Baidu ScholarGoogle Scholar
[157] K. J. Eskola, H. Paukkunen and C. A. Salgado,

EPS09: A new generation of nlo and lo nuclear parton distribution functions

. JHEP 04, 065 (2009). doi: 10.1088/1126-6708/2009/04/065
Baidu ScholarGoogle Scholar
[158] X. N. Wang,

Role of multiple mini-jets in high-energy hadronic reactions

. Phys. Rev. D 43, 104-112 (1991). doi: 10.1103/PhysRevD.43.104
Baidu ScholarGoogle Scholar
[159] L. D. McLerran and R. Venugopalan,

Computing quark and gluon distribution functions for very large nuclei

. Phys. Rev. D 49, 2233-2241 (1994). doi: 10.1103/PhysRevD.49.2233
Baidu ScholarGoogle Scholar
[160] C. Zhang, L. Zheng, S. Shi and Z. W. Lin,

Using local nuclear scaling of initial condition parameters to improve the system size dependence of transport model descriptions of nuclear collisions

. Phys. Rev. C. 104, 014908 (2021).
Baidu ScholarGoogle Scholar
[161] B. Muller and X. N. Wang,

Probing parton thermalization time with charm production

. Phys. Rev. Lett. 68, 2437-2439 (1992). doi: 10.1103/PhysRevLett.68.2437
Baidu ScholarGoogle Scholar
[162] Z. W. Lin and M. Gyulassy,

Open charm as a probe of preequilibrium dynamics in nuclear collisions

. Phys. Rev. C 51, 2177-2187 (1995). [erratum: Phys. Rev. C 52, 440 (1995)] doi: 10.1103/PhysRevC.52.440
Baidu ScholarGoogle Scholar
[163] X. Dong, Y. J. Lee and R. Rapp,

Open heavy-flavor production in heavy-ion collisions

. Ann. Rev. Nucl. Part. Sci. 69, 417-445 (2019). doi: 10.1146/annurev-nucl-101918-023806
Baidu ScholarGoogle Scholar
[164] S. K. Das, S. Plumari, S. Chatterjee, J. Alam, F. Scardina and V. Greco,

Directed flow of charm quarks as a witness of the initial strong magnetic field in ultra-relativistic heavy ion collisions

. Phys. Lett. B 768, 260-264 (2017). doi: 10.1016/j.physletb.2017.02.046
Baidu ScholarGoogle Scholar
[165] S. Chatterjee and P. Bożek,

Large directed flow of open charm mesons probes the three dimensional distribution of matter in heavy ion collisions

. Phys. Rev. Lett. 120, no.19, 192301 (2018). doi: 10.1103/PhysRevLett.120.192301
Baidu ScholarGoogle Scholar
[166] M. Nasim and S. Singha,

Directed flow of open charm in Au+Au collisions at sNN = 200 GeV using a quark coalescence model

. Phys. Rev. C 97, no.6, 064917 (2018). doi: 10.1103/PhysRevC.97.064917
Baidu ScholarGoogle Scholar
[167] M. He, R. J. Fries and R. Rapp,

Ds-meson as quantitative probe of diffusion and hadronization in nuclear collisions

. Phys. Rev. Lett. 110, no.11, 112301 (2013). doi: 10.1103/PhysRevLett.110.112301
Baidu ScholarGoogle Scholar
[168] K. Huggins and R. Rapp,

A T-Matrix calculation for in-medium heavy-quark gluon scattering

. Nucl. Phys. A 896, 24-45 (2012). doi: 10.1016/j.nuclphysa.2012.09.008
Baidu ScholarGoogle Scholar
[169] T. Lang, H. van Hees, J. Steinheimer, G. Inghirami and M. Bleicher,

Heavy quark transport in heavy ion collisions at energies available at the BNL Relativistic Heavy Ion Collider and at the CERN Large Hadron Collider within the UrQMD hybrid model

. Phys. Rev. C 93, no.1, 014901 (2016). doi: 10.1103/PhysRevC.93.014901
Baidu ScholarGoogle Scholar
[170] S. Cao et al.,

Toward the determination of heavy-quark transport coefficients in quark-gluon plasma

. Phys. Rev. C 99, no.5, 054907 (2019). doi: 10.1103/PhysRevC.99.054907
Baidu ScholarGoogle Scholar
[171] Y. Xu et al.,

Resolving discrepancies in the estimation of heavy quark transport coefficients in relativistic heavy-ion collisions

. Phys. Rev. C 99, no.1, 014902 (2019). doi: 10.1103/PhysRevC.99.014902
Baidu ScholarGoogle Scholar
[172] M. L. Mangano, P. Nason and G. Ridolfi,

Heavy quark correlations in hadron collisions at next-to-leading order

. Nucl. Phys. B 373, 295-345 (1992). doi: 10.1016/0550-3213(92)90435-E
Baidu ScholarGoogle Scholar
[173] B. A. Kniehl, G. Kramer, I. Schienbein and H. Spiesberger,

Inclusive D*+- production in p anti-p collisions with massive charm quarks

. Phys. Rev. D 71, 014018 (2005). doi: 10.1103/PhysRevD.71.014018
Baidu ScholarGoogle Scholar
[174] I. Helenius and H. Paukkunen,

Revisiting the D-meson hadroproduction in general-mass variable flavour number scheme

. JHEP 05, 196 (2018). doi: 10.1007/JHEP05(2018)196
Baidu ScholarGoogle Scholar
[175] M. Cacciari, P. Nason and R. Vogt,

QCD predictions for charm and bottom production at RHIC

. Phys. Rev. Lett. 95, 122001 (2005). doi: 10.1103/PhysRevLett.95.122001
Baidu ScholarGoogle Scholar
[176] M. Djordjevic and M. Djordjevic,

Predictions of heavy-flavor suppression at 5.1 TeV Pb + Pb collisions at the CERN Large Hadron Collider

. Phys. Rev. C 92, no.2, 024918 (2015). doi: 10.1103/PhysRevC.92.024918
Baidu ScholarGoogle Scholar
[177] J. Xu, J. Liao and M. Gyulassy,

Bridging soft-hard transport properties of quark-gluon plasmas with CUJET3.0

. JHEP 02, 169 (2016). doi: 10.1007/JHEP02(2016)169
Baidu ScholarGoogle Scholar
[178] A. Beraudo, A. De Pace, M. Monteno, M. Nardi and F. Prino,

Heavy flavors in heavy-ion collisions: quenching, flow and correlations

. Eur. Phys. J. C 75, no.3, 121 (2015). doi: 10.1140/epjc/s10052-015-3336-6
Baidu ScholarGoogle Scholar
[179] M. He, R. J. Fries and R. Rapp,

Heavy flavor at the large hadron collider in a strong coupling approach

. Phys. Lett. B 735, 445-450 (2014). doi: 10.1016/j.physletb.2014.05.050
Baidu ScholarGoogle Scholar
[180] S. Cao, G. Y. Qin and S. A. Bass,

Energy loss, hadronization and hadronic interactions of heavy flavors in relativistic heavy-ion collisions

. Phys. Rev. C 92, no.2, 024907 (2015). doi: 10.1103/PhysRevC.92.024907
Baidu ScholarGoogle Scholar
[181] J. Uphoff, O. Fochler, Z. Xu and C. Greiner,

Elastic and radiative heavy quark interactions in ultra-relativistic heavy-ion collisions

. J. Phys. G 42, no.11, 115106 (2015). doi: 10.1088/0954-3899/42/11/115106
Baidu ScholarGoogle Scholar
[182] S. Cao, T. Luo, G. Y. Qin and X. N. Wang,

Heavy and light flavor jet quenching at RHIC and LHC energies

. Phys. Lett. B 777, 255-259 (2018). doi: 10.1016/j.physletb.2017.12.023
Baidu ScholarGoogle Scholar
[183] M. Nahrgang, J. Aichelin, P. B. Gossiaux and K. Werner,

Influence of hadronic bound states above Tc on heavy-quark observables in Pb + Pb collisions at at the CERN Large Hadron Collider

. Phys. Rev. C 89, no.1, 014905 (2014). doi: 10.1103/PhysRevC.89.014905
Baidu ScholarGoogle Scholar
[184] T. Song, H. Berrehrah, D. Cabrera, W. Cassing and E. Bratkovskaya,

Charm production in Pb + Pb collisions at energies available at the CERN Large Hadron Collider

. Phys. Rev. C 93, no.3, 034906 (2016). doi: 10.1103/PhysRevC.93.034906
Baidu ScholarGoogle Scholar
[185] S. K. Das, F. Scardina, S. Plumari and V. Greco,

Toward a solution to the RAA and v2 puzzle for heavy quarks

. Phys. Lett. B 747, 260-264 (2015). doi: 10.1016/j.physletb.2015.06.003
Baidu ScholarGoogle Scholar
[186] S. Plumari, V. Minissale, S. K. Das, G. Coci and V. Greco,

Charmed Hadrons from Coalescence plus Fragmentation in relativistic nucleus-nucleus collisions at RHIC and LHC

. Eur. Phys. J. C 78, no.4, 348 (2018). doi: 10.1140/epjc/s10052-018-5828-7
Baidu ScholarGoogle Scholar
[187] B. Zhang, L. W. Chen and C. M. Ko,

Charm elliptic flow at RHIC

. Phys. Rev. C 72, 024906 (2005). doi: 10.1103/PhysRevC.72.024906
Baidu ScholarGoogle Scholar
[188] H. Li, Z. W. Lin and F. Wang,

Charm quarks are more hydrodynamic than light quarks in final-state elliptic flow

. Phys. Rev. C 99, no.4, 044911 (2019). doi: 10.1103/PhysRevC.99.044911
Baidu ScholarGoogle Scholar
[189] A. Kurkela, A. Mazeliauskas, J. F. Paquet, S. Schlichting and D. Teaney,

Matching the nonequilibrium initial stage of heavy ion collisions to hydrodynamics with QCD kinetic theory

. Phys. Rev. Lett. 122, no.12, 122302 (2019). doi: 10.1103/PhysRevLett.122.122302
Baidu ScholarGoogle Scholar
[190] A. Kurkela, A. Mazeliauskas, J. F. Paquet, S. Schlichting and D. Teaney,

Effective kinetic description of event-by-event pre-equilibrium dynamics in high-energy heavy-ion collisions

. Phys. Rev. C 99, no.3, 034910 (2019). doi: 10.1103/PhysRevC.99.034910
Baidu ScholarGoogle Scholar
[191] C. Lourenco and H. K. Wohri,

Heavy flavour hadro-production from fixed-target to collider energies

. Phys. Rept. 433, 127-180 (2006). doi: 10.1016/j.physrep.2006.05.005
Baidu ScholarGoogle Scholar
[192] A. Adare et al. [PHENIX],

Heavy quark production in p+p and energy loss and flow of heavy quarks in Au+Au collisions at sNN=200 GeV

. Phys. Rev. C 84, 044905 (2011). doi: 10.1103/PhysRevC.84.044905
Baidu ScholarGoogle Scholar
[193] L. Adamczyk et al. [STAR],

Measurements of D0 and D* production in p+p collisions at s=200 GeV

. Phys. Rev. D 86, 072013 (2012). doi: 10.1103/PhysRevD.86.072013
Baidu ScholarGoogle Scholar
[194] R. Aaij et al. [LHCb],

Prompt charm production in pp collisions at sqrt(s)=7 TeV

. Nucl. Phys. B 871, 1-20 (2013). doi: 10.1016/j.nuclphysb.2013.02.010
Baidu ScholarGoogle Scholar
[195] G. Aad et al. [ATLAS],

Measurement of D*±, D± and Ds± meson production cross sections in pp collisions at s=7 TeV with the ATLAS detector

. Nucl. Phys. B 907, 717-763 (2016). doi: 10.1016/j.nuclphysb.2016.04.032
Baidu ScholarGoogle Scholar
[196] B. Abelev et al. [ALICE],

Measurement of charm production at central rapidity in proton-proton collisions at s=2.76 TeV

. JHEP 07, 191 (2012). doi: 10.1007/JHEP07(2012)191
Baidu ScholarGoogle Scholar
[197] S. Acharya et al. [ALICE],

Measurement of D-meson production at mid-rapidity in pp collisions at s=7 TeV

. Eur. Phys. J. C 77, no.8, 550 (2017). doi: 10.1140/epjc/s10052-017-5090-4
Baidu ScholarGoogle Scholar
[198] J. Adam et al. [STAR],

Centrality and transverse momentum dependence of D0-meson production at mid-rapidity in Au+Au collisions at sNN=200 GeV

. Phys. Rev. C 99, no.3, 034908 (2019). doi: 10.1103/PhysRevC.99.034908
Baidu ScholarGoogle Scholar
[199] G. Xie [STAR],

Measurements of open charm hadron production in Au+Au Collisions at sNN = 200 GeV at STAR

. PoS HardProbes2018, 142 (2018). doi: 10.22323/1.345.0142
Baidu ScholarGoogle Scholar
[200] M. He and R. Rapp,

Charm-baryon production in proton-proton collisions

. Phys. Lett. B 795, 117-121 (2019). doi: 10.1016/j.physletb.2019.06.004
Baidu ScholarGoogle Scholar
[201] J. Adam et al. [STAR],

First measurement of Λc baryon production in Au+Au collisions at sNN = 200 GeV

. Phys. Rev. Lett. 124, no.17, 172301 (2020). doi: 10.1103/PhysRevLett.124.172301
Baidu ScholarGoogle Scholar
[202] Z. W. Lin and C. M. Ko,

A Model for J/psi absorption in hadronic matter

. Phys. Rev. C 62, 034903 (2000). doi: 10.1103/PhysRevC.62.034903
Baidu ScholarGoogle Scholar
[203] Z. W. Lin, C. M. Ko and B. Zhang,

Hadronic scattering of charm mesons

. Phys. Rev. C 61, 024904 (2000). doi: 10.1103/PhysRevC.61.024904
Baidu ScholarGoogle Scholar
[204] Z. W. Lin, T. G. Di and C. M. Ko,

Charm meson scattering cross-sections by pion and rho meson

. Nucl. Phys. A 689, 965-979 (2001). doi: 10.1016/S0375-9474(00)00611-4
Baidu ScholarGoogle Scholar
[205] A. Sibirtsev, K. Tsushima and A. W. Thomas,

Charmonium absorption by nucleons

. Phys. Rev. C 63, 044906 (2001). doi: 10.1103/PhysRevC.63.044906
Baidu ScholarGoogle Scholar
[206] C. Bierlich, T. Sjöstrand and M. Utheim,

Hadronic rescattering in pA and AA collisions

. Eur. Phys. J. A 57, no.7, 227 (2021). doi: 10.1140/epja/s10050-021-00543-3
Baidu ScholarGoogle Scholar
[207] B. Andersson and P. A. Henning,

On the dynamics of a color rope: The Fragmentation of interacting strings and the longitudinal distributions

. Nucl. Phys. B 355, 82-105 (1991). doi: 10.1016/0550-3213(91)90303-F
Baidu ScholarGoogle Scholar
[208] C. Bierlich and J. R. Christiansen,

Effects of color reconnection on hadron flavor observables

. Phys. Rev. D 92, no.9, 094010 (2015). doi: 10.1103/PhysRevD.92.094010
Baidu ScholarGoogle Scholar
[209] C. Bierlich, G. Gustafson, L. Lönnblad and A. Tarasov,

Effects of overlapping strings in pp collisions

. JHEP 03, 148 (2015). doi: 10.1007/JHEP03(2015)148
Baidu ScholarGoogle Scholar
[210] N. Fischer and T. Sjöstrand,

Thermodynamical string fragmentation

. JHEP 01, 140 (2017). doi: 10.1007/JHEP01(2017)140
Baidu ScholarGoogle Scholar
[211] C. Bierlich,

Rope hadronization and strange particle production

. EPJ Web Conf. 171, 14003 (2018). doi: 10.1051/epjconf/201817114003
Baidu ScholarGoogle Scholar
[212] T. S. Biro, H. B. Nielsen and J. Knoll,

Color rope model for extreme relativistic heavy ion collisions

. Nucl. Phys. B 245, 449-468 (1984). doi: 10.1016/0550-3213(84)90441-3
Baidu ScholarGoogle Scholar
[213] A. Tai and B. H. Sa,

Increase of effective string tension and production of strange particles

. Phys. Lett. B 409, 393 (1997). doi: 10.1016/S0370-2693(97)00793-4
Baidu ScholarGoogle Scholar
[214] J. S. Moreland, J. E. Bernhard and S. A. Bass,

Alternative ansatz to wounded nucleon and binary collision scaling in high-energy nuclear collisions

. Phys. Rev. C 92, no.1, 011901 (2015). doi: 10.1103/PhysRevC.92.011901
Baidu ScholarGoogle Scholar
[215] J. E. Bernhard, J. S. Moreland, S. A. Bass, J. Liu and U. Heinz,

Applying Bayesian parameter estimation to relativistic heavy-ion collisions: simultaneous characterization of the initial state and quark-gluon plasma medium

. Phys. Rev. C 94, no.2, 024907 (2016). doi: 10.1103/PhysRevC.94.024907
Baidu ScholarGoogle Scholar
[216] L. Adamczyk et al. [STAR],

Bulk properties of the medium produced in relativistic heavy-ion collisions from the Beam Energy Scan program

. Phys. Rev. C 96, no.4, 044904 (2017). doi: 10.1103/PhysRevC.96.044904
Baidu ScholarGoogle Scholar
[217] B. B. Back et al. [PHOBOS],

Centrality dependence of charged hadron transverse momentum spectra in Au + Au collisions from s(NN)**(1/2) = 62.4-GeV to 200-GeV

. Phys. Rev. Lett. 94, 082304 (2005). doi: 10.1103/PhysRevLett.94.082304
Baidu ScholarGoogle Scholar
[218] J. Adams et al. [STAR],

Transverse momentum and collision energy dependence of high p(T) hadron suppression in Au+Au collisions at ultrarelativistic energies

. Phys. Rev. Lett. 91, 172302 (2003). doi: 10.1103/PhysRevLett.91.172302
Baidu ScholarGoogle Scholar
[219] B. Abelev et al. [ALICE],

Centrality dependence of charged particle production at large transverse momentum in Pb-Pb collisions at sNN=2.76 TeV

. Phys. Lett. B 720, 52-62 (2013). doi: 10.1016/j.physletb.2013.01.051
Baidu ScholarGoogle Scholar
[220] S. Acharya et al. [ALICE],

Transverse momentum spectra and nuclear modification factors of charged particles in pp, p-Pb and Pb-Pb collisions at the LHC

. JHEP 11, 013 (2018). doi: 10.1007/JHEP11(2018)013
Baidu ScholarGoogle Scholar
[221] A. Adare et al. [PHENIX],

Transverse energy production and charged-particle multiplicity at midrapidity in various systems from sNN=7.7 to 200 GeV

. Phys. Rev. C 93, no.2, 024901 (2016). doi: 10.1103/PhysRevC.93.024901
Baidu ScholarGoogle Scholar
[222] J. Adam et al. [ALICE],

Centrality dependence of the pseudorapidity density distribution for charged particles in Pb-Pb collisions at sNN=5.02 TeV

. Phys. Lett. B 772, 567-577 (2017). doi: 10.1016/j.physletb.2017.07.017
Baidu ScholarGoogle Scholar
[223] S. Acharya et al. [ALICE],

Transverse momentum spectra and nuclear modification factors of charged particles in Xe-Xe collisions at sNN = 5.44 TeV

. Phys. Lett. B 788, 166-179 (2019). doi: 10.1016/j.physletb.2018.10.052
Baidu ScholarGoogle Scholar
[224] S. Acharya et al. [ALICE],

Centrality and pseudorapidity dependence of the charged-particle multiplicity density in Xe–Xe collisions at sNN =5.44TeV

. Phys. Lett. B 790, 35-48 (2019). doi: 10.1016/j.physletb.2018.12.048
Baidu ScholarGoogle Scholar
[225] B. Alver et al. [PHOBOS],

System size and centrality dependence of charged hadron transverse momentum spectra in Au + Au and Cu + Cu collisions at s(NN)**(1/2) = 62.4-GeV and 200-GeV

. Phys. Rev. Lett. 96, 212301 (2006). doi: 10.1103/PhysRevLett.96.212301
Baidu ScholarGoogle Scholar
[226] J. Adam et al. [ALICE],

Centrality dependence of particle production in p-Pb collisions at sNN = 5.02 TeV

. Phys. Rev. C 91, no.6, 064905 (2015). doi: 10.1103/PhysRevC.91.064905
Baidu ScholarGoogle Scholar
[227] P. Möller, A. J. Sierk, T. Ichikawa and H. Sagawa,

Nuclear ground-state masses and deformations: FRDM(2012)

. Atom. Data Nucl. Data Tabl. 109-110, 1-204 (2016). doi: 10.1016/j.adt.2015.10.002
Baidu ScholarGoogle Scholar
[228] T. Sjöstrand, S. Ask, J. R. Christiansen, R. Corke, N. Desai, P. Ilten, S. Mrenna, S. Prestel, C. O. Rasmussen and P. Z. Skands,

An introduction to PYTHIA 8.2

. Comput. Phys. Commun. 191, 159-177 (2015). doi: 10.1016/j.cpc.2015.01.024
Baidu ScholarGoogle Scholar
[229] C. Bierlich, G. Gustafson, L. Lönnblad and H. Shah,

The Angantyr model for heavy-ion collisions in PYTHIA8

. JHEP 10, 134 (2018). doi: 10.1007/JHEP10(2018)134
Baidu ScholarGoogle Scholar
[230] Y. Kanakubo, Y. Tachibana and T. Hirano,

Unified description of hadron yield ratios from dynamical core-corona initialization

. Phys. Rev. C 101, no.2, 024912 (2020). doi: 10.1103/PhysRevC.101.024912
Baidu ScholarGoogle Scholar
[231] G. Papp, G. Gábor Barnaföldi, G. Bíró, M. Gyulassy, S. Harangozó, G. Ma, P. Lévai, X. N. Wang and B. W. Zhang,

First results with HIJING++ on high-energy heavy ion collisions

. arXiv:1805.02635 [hep-ph].
Baidu ScholarGoogle Scholar
[232] J. H. Putschke et al.,

The JETSCAPE framework

. arXiv:1903.07706 [nucl-th].
Baidu ScholarGoogle Scholar
[233] L. Zheng, G. H. Zhang, Y. F. Liu, Z. W. Lin, Q. Y. Shou and Z. B. Yin,

Investigating high energy proton proton collisions with a multi-phase transport model approach based on PYTHIA8 initial conditions

. arXiv:2104.05998 [hep-ph].
Baidu ScholarGoogle Scholar
[234] B. Schenke and R. Venugopalan,

Eccentric protons? Sensitivity of flow to system size and shape in p+p, p+Pb and Pb+Pb collisions

. Phys. Rev. Lett. 113, 102301 (2014). doi: 10.1103/PhysRevLett.113.102301
Baidu ScholarGoogle Scholar
[235] H. Mäntysaari and B. Schenke,

Evidence of strong proton shape fluctuations from incoherent diffraction

. Phys. Rev. Lett. 117, no.5, 052301 (2016). doi: 10.1103/PhysRevLett.117.052301
Baidu ScholarGoogle Scholar
[236] K. Welsh, J. Singer and U. W. Heinz,

Initial state fluctuations in collisions between light and heavy ions

. Phys. Rev. C 94, no.2, 024919 (2016). doi: 10.1103/PhysRevC.94.024919
Baidu ScholarGoogle Scholar
[237] H. Mäntysaari, B. Schenke, C. Shen and P. Tribedy,

Imprints of fluctuating proton shapes on flow in proton-lead collisions at the LHC

. Phys. Lett. B 772, 681-686 (2017). doi: 10.1016/j.physletb.2017.07.038
Baidu ScholarGoogle Scholar
[238] D. d’Enterria, G. K. Eyyubova, V. L. Korotkikh, I. P. Lokhtin, S. V. Petrushanko, L. I. Sarycheva and A. M. Snigirev,

Estimates of hadron azimuthal anisotropy from multiparton interactions in proton-proton collisions at s=14 TeV

. Eur. Phys. J. C 66, 173-185 (2010). doi: 10.1140/epjc/s10052-009-1232-7
Baidu ScholarGoogle Scholar
[239] S. Eremin and S. Voloshin,

Nucleon participants or quark participants?

. Phys. Rev. C 67, 064905 (2003). doi: 10.1103/PhysRevC.67.064905
Baidu ScholarGoogle Scholar
[240] L. Zheng and Z. Yin,

A systematic study of the initial state in heavy ion collisions based on the quark participant assumption

. Eur. Phys. J. A 52, 45 (2016). doi: 10.1140/epja/i2016-16045-x
Baidu ScholarGoogle Scholar
[241] C. Loizides,

Glauber modeling of high-energy nuclear collisions at the subnucleon level

. Phys. Rev. C 94, no.2, 024914 (2016). doi: 10.1103/PhysRevC.94.024914
Baidu ScholarGoogle Scholar
[242] P. Bożek, W. Broniowski and M. Rybczyński,

Wounded quarks in A+A, p+A, and p+p collisions

. Phys. Rev. C 94, no.1, 014902 (2016). doi: 10.1103/PhysRevC.94.014902
Baidu ScholarGoogle Scholar
[243] P. Bożek, W. Broniowski, M. Rybczynski and G. Stefanek,

GLISSANDO 3: GLauber Initial-State Simulation AND mOre, ver. 3

. Comput. Phys. Commun. 245, 106850 (2019). doi: 10.1016/j.cpc.2019.07.014
Baidu ScholarGoogle Scholar
[244] D. d’Enterria and C. Loizides,

Progress in the Glauber model at collider energies

. arXiv:2011.14909 [hep-ph].
Baidu ScholarGoogle Scholar
[245] S. Acharya et al. [ALICE],

Multiplicity dependence of π , K, and p production in pp collisions at s=13 TeV

. Eur. Phys. J. C 80, no.8, 693 (2020). doi: 10.1140/epjc/s10052-020-8125-1
Baidu ScholarGoogle Scholar
[246] J. L. Nagle, R. Belmont, K. Hill, J. Orjuela Koop, D. V. Perepelitsa, P. Yin, Z. W. Lin and D. McGlinchey,

Minimal conditions for collectivity in e+e- and p+p collisions

. Phys. Rev. C 97, no.2, 024909 (2018). doi: 10.1103/PhysRevC.97.024909
Baidu ScholarGoogle Scholar
[247] V. Khachatryan et al. [CMS],

Evidence for collectivity in pp collisions at the LHC

. Phys. Lett. B 765, 193-220 (2017). doi: 10.1016/j.physletb.2016.12.009
Baidu ScholarGoogle Scholar
[248] T. Sjöstrand and M. Utheim,

A framework for hadronic rescattering in pp collisions

. Eur. Phys. J. C 80, no.10, 907 (2020). doi: 10.1140/epjc/s10052-020-8399-3
Baidu ScholarGoogle Scholar
[249] S. Ferreres-Solé and T. Sjöstrand,

The space–time structure of hadronization in the Lund model

. Eur. Phys. J. C 78, no.11, 983 (2018). doi: 10.1140/epjc/s10052-018-6459-8
Baidu ScholarGoogle Scholar
[250] A. V. da Silva, W. M. Serenone, D. Dobrigkeit Chinellato, J. Takahashi and C. Bierlich,

Studies of heavy-ion collisions using PYTHIA Angantyr and UrQMD

. arXiv:2002.10236 [hep-ph].
Baidu ScholarGoogle Scholar
[251] T. Kodama, S. B. Duarte, K. C. Chung, R. Donangelo and R. A. M. S. Nazareth,

Causality and relativistic effects in intranuclear cascade calculations

. Phys. Rev. C 29, 2146-2152 (1984). doi: 10.1103/PhysRevC.29.2146
Baidu ScholarGoogle Scholar
[252] G. Kortemeyer, W. Bauer, K. Haglin, J. Murray and S. Pratt,

Causality violations in cascade models of nuclear collisions

. Phys. Rev. C 52, 2714-2724 (1995). doi: 10.1103/PhysRevC.52.2714
Baidu ScholarGoogle Scholar
[253] B. Zhang and Y. Pang,

Frame dependence of parton cascade results

. Phys. Rev. C 56, 2185-2190 (1997). doi: 10.1103/PhysRevC.56.2185
Baidu ScholarGoogle Scholar
[254] B. Zhang, M. Gyulassy and Y. Pang,

Equation of state and collision rate tests of parton cascade models

. Phys. Rev. C 58, 1175-1182 (1998). doi: 10.1103/PhysRevC.58.1175
Baidu ScholarGoogle Scholar
[255] S. Cheng, S. Pratt, P. Csizmadia, Y. Nara, D. Molnar, M. Gyulassy, S. E. Vance and B. Zhang,

The Effect of finite range interactions in classical transport theory

. Phys. Rev. C 65, 024901 (2002). doi: 10.1103/PhysRevC.65.024901
Baidu ScholarGoogle Scholar
[256] C. Y. Wong,

Dynamics of nuclear fluid. VIII. Time-dependent Hartree-Fock approximation from a classical point of view

. Phys. Rev. C 25, 1460-1475 (1982). doi: 10.1103/PhysRevC.25.1460
Baidu ScholarGoogle Scholar
[257] G. Welke, R. Malfliet, C. Gregoire, M. Prakash and E. Suraud,

Collisional relaxation in simulations of heavy-ion collisions using Boltzmann-type equations

. Phys. Rev. C 40, 2611-2620 (1989). doi: 10.1103/PhysRevC.40.2611
Baidu ScholarGoogle Scholar
[258] Y. Pang, https://karman.physics.purdue.edu/OSCAR-old/rhic/gcp/proc.html;

General Cascade Program

, in Proceedings of RHIC’96, CU-TP-815 (1997).
Baidu ScholarGoogle Scholar
[259] D. Molnar and M. Gyulassy,

New solutions to covariant nonequilibrium dynamics

. Phys. Rev. C 62, 054907 (2000). doi: 10.1103/PhysRevC.62.054907
Baidu ScholarGoogle Scholar
[260] D. Molnar and P. Huovinen,

Dissipation and elliptic flow at RHIC

. Phys. Rev. Lett. 94, 012302 (2005). doi: 10.1103/PhysRevLett.94.012302
Baidu ScholarGoogle Scholar
[261] X. L. Zhao, G. L. Ma, Y. G. Ma and Z. W. Lin,

Validation and improvement of the ZPC parton cascade inside a box

. Phys. Rev. C 102, no.2, 024904 (2020). doi: 10.1103/PhysRevC.102.024904
Baidu ScholarGoogle Scholar
[262] T. Schäfer and D. Teaney,

Nearly perfect fluidity: from cold atomic gases to hot quark gluon plasmas

. Rept. Prog. Phys. 72, 126001 (2009). doi: 10.1088/0034-4885/72/12/126001
Baidu ScholarGoogle Scholar
[263] M. S. Green,

Markoff random processes and the statistical mechanics of time-dependent phenomena. II. Irreversible processes in fluids

. J. Chem. Phys. 22, 398 (1954).
Baidu ScholarGoogle Scholar
[264] R. Kubo,

Statistical-mechanical theory of irreversible processes. I. General theory and simple applications to magnetic and conduction problems

. J. Phys. Soc. Jpn. 12, 570 (1957).
Baidu ScholarGoogle Scholar
[265] A. Muronga,

Shear viscosity coefficient from microscopic models

. Phys. Rev. C 69, 044901 (2004). doi: 10.1103/PhysRevC.69.044901
Baidu ScholarGoogle Scholar
[266] N. Demir and S. A. Bass,

Shear-viscosity to entropy-density ratio of a relativistic hadron gas

. Phys. Rev. Lett. 102, 172302 (2009). doi: 10.1103/PhysRevLett.102.172302
Baidu ScholarGoogle Scholar
[267] J., III Fuini, N. S. Demir, D. K. Srivastava and S. A. Bass,

Shear viscosity in a perturbative quark-gluon-plasma

. J. Phys. G 38, 015004 (2011). doi: 10.1088/0954-3899/38/1/015004
Baidu ScholarGoogle Scholar
[268] C. Wesp, A. El, F. Reining, Z. Xu, I. Bouras and C. Greiner,

Calculation of shear viscosity using Green-Kubo relations within a parton cascade

. Phys. Rev. C 84, 054911 (2011). doi: 10.1103/PhysRevC.84.054911
Baidu ScholarGoogle Scholar
[269] S. X. Li, D. Q. Fang, Y. G. Ma and C. L. Zhou,

Shear viscosity to entropy density ratio in the Boltzmann-Uehling-Uhlenbeck model

. Phys. Rev. C 84, 024607 (2011). doi: 10.1103/PhysRevC.84.024607
Baidu ScholarGoogle Scholar
[270] M. R. Haque, Z. W. Lin and B. Mohanty,

Multiplicity, average transverse momentum and azimuthal anisotropy in U+U collisions at sNN = 200 GeV using AMPT model

. Phys. Rev. C 85, 034905 (2012). doi: 10.1103/PhysRevC.85.034905
Baidu ScholarGoogle Scholar
[271] H. J. Xu, X. Wang, H. Li, J. Zhao, Z. W. Lin, C. Shen and F. Wang,

Importance of isobar density distributions on the chiral magnetic effect search

. Phys. Rev. Lett. 121, no.2, 022301 (2018). doi: 10.1103/PhysRevLett.121.022301
Baidu ScholarGoogle Scholar
[272] H. j. Xu, J. Zhao, X. Wang, H. Li, Z. W. Lin, C. Shen and F. Wang,

Varying the chiral magnetic effect relative to flow in a single nucleus-nucleus collision

. Chin. Phys. C 42, no.8, 084103 (2018). doi: 10.1088/1674-1137/42/8/084103
Baidu ScholarGoogle Scholar
[273] H. Li, H. j. Xu, J. Zhao, Z. W. Lin, H. Zhang, X. Wang, C. Shen and F. Wang,

Multiphase transport model predictions of isobaric collisions with nuclear structure from density functional theory

. Phys. Rev. C 98, no.5, 054907 (2018). doi: 10.1103/PhysRevC.98.054907
Baidu ScholarGoogle Scholar
[274] J. Xu, L. W. Chen, C. M. Ko and Z. W. Lin,

Effects of hadronic potentials on elliptic flows in relativistic heavy ion collisions

. Phys. Rev. C 85, 041901 (2012). doi: 10.1103/PhysRevC.85.041901
Baidu ScholarGoogle Scholar
[275] J. Xu, T. Song, C. M. Ko and F. Li,

Elliptic flow splitting as a probe of the QCD phase structure at finite baryon chemical potential

. Phys. Rev. Lett. 112, 012301 (2014). doi: 10.1103/PhysRevLett.112.012301
Baidu ScholarGoogle Scholar
[276] Z. W. Lin,

Recent developments of a multi-phase transport model

. Acta Phys. Polon. Supp. 7, no.1, 191-197 (2014). doi: 10.5506/APhysPolBSupp.7.191
Baidu ScholarGoogle Scholar
[277] Z. W. Lin and G. L. Ma, unpublished.
[278] A. H. Tang,

Probe chiral magnetic effect with signed balance function

. Chin. Phys. C 44, no.5, 054101 (2020). doi: 10.1088/1674-1137/44/5/054101
Baidu ScholarGoogle Scholar
[279] S. Choudhury, G. Wang, W. He, Y. Hu and H. Z. Huang,

Background evaluations for the chiral magnetic effect with normalized correlators using a multiphase transport model

. Eur. Phys. J. C 80, no.5, 383 (2020). doi: 10.1140/epjc/s10052-020-7928-4
Baidu ScholarGoogle Scholar
[280] G. C. Yong, Z. G. Xiao, Y. Gao, and Z. W. Lin,

Double strangeness Ξ− production as a probe of nuclear equation of state at high densities

. Phys. Lett. B 820, 136521 (2021). doi: 10.1016/j.physletb.2021.136521
Baidu ScholarGoogle Scholar