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

Strange hadron production in a quark combination model in Au+Au collisions at energies available at the BNL Relativistic Heavy Ion Collider

Jun Song Department of Physics, Jining University, Shandong 273155, China    Xiao-feng Wang School of Physics and Physical Engineering, Qufu Normal University, Shandong 273165, China    Hai-hong Li Department of Physics, Jining University, Shandong 273155, China    Rui-qin Wang School of Physics and Physical Engineering, Qufu Normal University, Shandong 273165, China    Feng-lan Shao [email protected] School of Physics and Physical Engineering, Qufu Normal University, Shandong 273165, China
Abstract

We apply a quark combination model to study yield densities and transverse momentum (pTp_{T}) spectra of strange (anti-)hadrons at mid-rapidity in central Au+Au collisions at sNN=\sqrt{s_{NN}}= 7.7, 11.5, 19.6, 27, 39 and 200 GeV. We show that the experimental data for pTp_{T} spectra of (anti-)hadrons in these collisions can be systematically described by the equal velocity combination of constituent quarks and antiquarks at hadronization. We obtain the pTp_{T} spectra of quarks and antiquarks at hadronization and study their collision energy dependence. We also reproduce the yield densities of hadrons and anti-hadrons. In particular, we demonstrate that the yield ratios of anti-hadrons to hadrons K/K+K^{-}/K^{+}, p¯/p\bar{p}/p, Λ¯/Λ\bar{\Lambda}/\Lambda, Ξ¯+/Ξ\bar{\Xi}^{+}/\Xi^{-} and Ω¯+/Ω\bar{\Omega}^{+}/\Omega^{-} simply correlate with each other and their experimental data except Ω¯+/Ω\bar{\Omega}^{+}/\Omega^{-} at sNN=\sqrt{s_{NN}}= 7.7 GeV are systematically described by the model. These results suggest that the equal velocity combination mechanism for quarks and antiquarks at hadronization plays an important role for the production of these long-lived hadrons in Au+Au collisions at low RHIC energies (sNN\sqrt{s_{NN}}\geq 11.5 GeV).

I Introduction

Hadron production from final state partons in high energy collisions is a complex Quantum Chromodynamics (QCD) process. Due to the difficulty of non-perturbative QCD, phenomenological mechanisms and models have to be applied to describe the production of hadrons at hadronization Andersson et al. (1983); Webber (1984); Das and Hwa (1977); Anisovich and Shekhter (1973); Bjorken and Farrar (1974); Xie and Liu (1988); Becattini (1996); Braun-Munzinger et al. (1996). In relativistic heavy-ion collisions at high RHIC and LHC energies, quark-gluon plasma (QGP) is created in early stage of collisions and the hadronization of QGP can be microscopically described by the quark (re-)combination/coalescence mechanism Greco et al. (2003); Fries et al. (2003); Molnar and Voloshin (2003); Lin and Molnar (2003); Shao et al. (2005); Chen and Ko (2006). The enhanced ratio of baryon to meson and number of constituent quark scaling for elliptic flows of hadrons at the intermediate transverse momentum (pTp_{T}) are typical experimental signals for quark combination mechanism at hadronization and have been widely observed in relativistic heavy-ion collisions Adare et al. (2007); Adamczyk et al. (2016a); Abelev et al. (2015); Acharya et al. (2018); Adamczyk et al. (2013); Adcox et al. (2002a); Abelev et al. (2006, 2013).

In our recent studies in high-multiplicity events in pppp and pp-Pb collisions at LHC energies where the mini-QGP is possibly created and re-scattering of hadrons is weak, we found an interesting quark number scaling property for pTp_{T} spectra of hadrons at mid-rapidity Song et al. (2017); Zhang et al. (2020). This scaling property is a direct consequence of equal velocity combination (EVC) of constituent quarks and antiquarks at hadronization. Our studies showed that the EVC of up/down, strange and charm quarks can provide a good and systematic description on pTp_{T} spectra of light, strange and charm hadrons in ground-state in pppp and pp-Pb collisions at LHC energies Song et al. (2017); Gou et al. (2017); Zhang et al. (2020); Song et al. (2018); Li et al. (2018). In latest work Song et al. (2020), we further found that the experimental data for pTp_{T} spectra of Ω\Omega and ϕ\phi at mid-rapidity in heavy-ion collisions in a broad collision energy range (sNN=11.52760\sqrt{s_{NN}}=11.5-2760 GeV) also satisfy the quark number scaling property. This is a clear signal of EVC at hadronization even in heavy-ion collisions. Therefore, it is interesting to systematically test this mechanism by production of more hadron species in relativistic heavy-ion collisions.

Recently, STAR collaboration reported their precise experimental data for the production of strange hadrons in Au+Au collisions at sNN=\sqrt{s_{NN}}=7.7-39 GeV Adam et al. (2020). This provides us a good opportunity to systematically study the EVC mechanism of hadron production in these collisions. In this paper, we apply a quark combination model with EVC to carry out a systematic study on yield densities and pTp_{T} spectra of strange hadrons in Au+Au collisions at sNN=\sqrt{s_{NN}}= 7.7, 11.5, 19.6, 27, 39 and 200 GeV. We put particular emphasis on the self-consistency of the model in explaining the experimental data for different kinds of hadrons and on the regularity in multi-hadron production correlations which is sensitive to hadronization mechanism. Taking advantage of precise data for strange hadrons, we also study the strangeness neutralization in the midrapidity region in these collisions. We extract quark pTp_{T} distributions at hadronization from data of hadrons and study properties of the relative abundance for strange quarks as the function of collision energy. Furthermore, we discuss the key physics in current quark combination model which are responsible for explaining successfully experimental data of hadronic pTp_{T} spectra and yields, and discuss the creation of QGP or the deconfinement at low RHIC energies.

The paper is organized as follows. In Sec. II, we introduce a quark combination model with equal velocity combination approximation. In Sec. III, we study the strangeness neutralization in mid-rapidity region. In Sec. IV, we show results of pTp_{T} spectra for hadrons and compare them with experimental data. In Sec. V, we study the split in yield between hadrons and their antiparticles. In Sec. VI, we study properties for the obtained numbers and pTp_{T} spectra of quarks at hadronization at different collision energies. The summary and discussion are given at last in Sec. VII.

II A quark combination model with EVC

The quark combination is one of phenomenological mechanisms for hadron production at hadronization. The basic idea of quark combination was firstly proposed in 1970s Anisovich and Shekhter (1973); Bjorken and Farrar (1974); Das and Hwa (1977) and has many successful applications in high energy reactions Xie and Liu (1988); Buschbeck et al. (1980); Liang and Xie (1991); Braaten et al. (2002). In relativistic heavy-ion collisions, quark combination mechanism is often used to describe the hadron production at QGP hadronization Zimányi et al. (2000); Hwa and Yang (2003); Greco et al. (2003); Fries et al. (2003); Molnar and Voloshin (2003); Shao et al. (2005); Chen and Ko (2006). Recently, we found that quark combination can also well explain experimental data of hadron production in high-multiplicity pppp and pp-Pb collisions at LHC energies Song et al. (2017); Gou et al. (2017); Shao et al. (2017); Zhang et al. (2020).

In this paper, inspired by the quark number scaling property of hadronic pTp_{T} spectra Song et al. (2017); Zhang et al. (2020); Song et al. (2020), we adopt a specific version of quark combination model Song et al. (2017); Gou et al. (2017). This model assumes the combination of constituent quarks and antiquarks with equal velocity to form baryons and mesons at hadronization. The unknown non-perturbative dynamics at hadronization are parameterized and their values are assumed to be stable in high energy reactions and are fixed by experimental data. This model is essentially a statistical model based on the constituent quark degrees of freedom at hadronization and the constituent quark structure of hadrons. It is different from the popular re-combination/coalescence models which adopt the Wigner wave function method under instantaneous hadronization approximation Fries et al. (2003); Greco et al. (2003). The brief description of the model is as follows.

We start from the general formula for the production of the baryon BjB_{j} composed of q1q2q3q_{1}q_{2}q_{3} and the production of the meson MjM_{j} composed of q1q¯2q_{1}\bar{q}_{2} in quark combination mechanism

fBj(pB)\displaystyle f_{B_{j}}(p_{B}) =𝑑p1𝑑p2𝑑p3Bj(p1,p2,p3;pB)\displaystyle=\int dp_{1}dp_{2}dp_{3}{\cal R}_{B_{j}}(p_{1},p_{2},p_{3};p_{B}) (1)
×fq1q2q3(p1,p2,p3),\displaystyle\,\,\,\,\,\,\,\,\times f_{q_{1}q_{2}q_{3}}(p_{1},p_{2},p_{3}),
fMj(pM)\displaystyle f_{M_{j}}(p_{M}) =𝑑p1𝑑p2Mj(p1,p2;pM)fq1q¯2(p1,p2).\displaystyle=\int dp_{1}dp_{2}{\cal R}_{M_{j}}(p_{1},p_{2};p_{M})f_{q_{1}\bar{q}_{2}}(p_{1},p_{2}). (2)

Here fq1q2q3(p1,p2,p3)f_{q_{1}q_{2}q_{3}}(p_{1},p_{2},p_{3}) is the joint momentum distribution for q1q_{1}, q2q_{2} and q3q_{3}. The combination kernel function Bj(p1,p2,p3;pB){\cal R}_{B_{j}}(p_{1},p_{2},p_{3};p_{B}) denotes the probability density for a given q1q2q3q_{1}q_{2}q_{3} with momenta p1,p2p_{1},p_{2} and p3p_{3} to combine into a baryon BjB_{j} with momentum pBp_{B}. It is similar for mesons. We emphasize that Eqs. (1) and (2) are generally suitable for hadron production in momentum space of any dimension. In this paper, we study the one-dimensional transverse momentum (pTp_{T}) distribution of hadrons at midrapidity y=0y=0. In this case, pip_{i} simply denotes pT,ip_{T,i} and the distribution function fhf_{h}(p)\left(p\right) denotes the dNh/dpTdN_{h}/dp_{T} at midrapidity.

The combination functions Bj(p1,p2,p3;pB){\cal R}_{B_{j}}(p_{1},p_{2},p_{3};p_{B}) and Mj(p1,p2;pM){\cal R}_{M_{j}}(p_{1},p_{2};p_{M}) contain the key information of combination dynamics which is not clear at present due to the non-perturbative difficulty of hadronization. In our recent works Song et al. (2017); Zhang et al. (2020); Song et al. (2020), we observed an interesting quark number scaling property for the pTp_{T} spectra of hadrons in high-multiplicity pppp and pp-Pb collisions as well as in heavy-ion collisions. This scaling property supports the combination of constituent quarks and antiquarks with equal velocity. This suggests an effective form for the combination kernel functions, i.e.,

Bj(p1,p2,p3;pB)\displaystyle{\cal R}_{B_{j}}(p_{1},p_{2},p_{3};p_{B}) =κBji=13δ(pixipB),\displaystyle=\kappa_{B_{j}}\prod_{i=1}^{3}\delta(p_{i}-x_{i}p_{B}), (3)
Mj(p1,p2;pM)\displaystyle{\cal R}_{M_{j}}(p_{1},p_{2};p_{M}) =κMji=12δ(pixipM).\displaystyle=\kappa_{M_{j}}\prod_{i=1}^{2}\delta(p_{i}-x_{i}p_{M}). (4)

Here, κBj\kappa_{B_{j}} and κMj\kappa_{M_{j}} are coefficients independent of momentum. Momentum fraction xix_{i} is determined by the masses of constituent quarks because pi=miγβmip_{i}=m_{i}\gamma\beta\propto m_{i}. Specifically, we have xi=mi/(m1+m2+m3)x_{i}=m_{i}/(m_{1}+m_{2}+m_{3}) for baryon BjB_{j} with x1+x2+x3=1x_{1}+x_{2}+x_{3}=1 and xi=mi/(m1+m2)x_{i}=m_{i}/(m_{1}+m_{2}) for meson MjM_{j} with x1+x2=1x_{1}+x_{2}=1. mim_{i} is the constituent mass for quark of flavor ii. Because the mass of the formed hadron under Eqs. (3) and (4) is the sum of these of constituent quarks, we take contituent masses mu=md=0.3m_{u}=m_{d}=0.3 GeV and ms=0.5m_{s}=0.5 GeV in order to properly describe the production of baryons and vector mesons studied in this paper.

The joint momentum distributions fq1q2q3(p1,p2,p3)f_{q_{1}q_{2}q_{3}}(p_{1},p_{2},p_{3}) and fq1q¯2(p1,p2)f_{q_{1}\bar{q}_{2}}(p_{1},p_{2}) generally contain the correlation term caused by, for example, the collective flow formed in system evolution before hadronization in heavy-ion collisions. In order to obtain analytical and simple expressions for fBj(pB)f_{B_{j}}(p_{B}) and fMj(pM)f_{M_{j}}(p_{M}), we take the independent distribution approximation

fq1q2q3(p1,p2,p3)\displaystyle f_{q_{1}q_{2}q_{3}}(p_{1},p_{2},p_{3}) =fq1(p1)fq2(p2)fq3(p3),\displaystyle=f_{q_{1}}(p_{1})f_{q_{2}}(p_{2})f_{q_{3}}(p_{3}), (5)
fq1q¯2(p1,p2)\displaystyle f_{q_{1}\bar{q}_{2}}(p_{1},p_{2}) =fq1(p1)fq¯2(p2).\displaystyle=f_{q_{1}}(p_{1})f_{\bar{q}_{2}}(p_{2}). (6)

Substituting Eqs. (3)-(6) into Eqs. (1) and (2), we obtain

fBj(pB)\displaystyle f_{B_{j}}(p_{B}) =κBjfq1(x1pB)fq2(x2pB)fq3(x3pB),\displaystyle=\kappa_{B_{j}}f_{q_{1}}(x_{1}p_{B})f_{q_{2}}(x_{2}p_{B})f_{q_{3}}(x_{3}p_{B}), (7)
fMj(pM)\displaystyle f_{M_{j}}(p_{M}) =κMjfq1(x1pM)fq¯2(x2pM).\displaystyle=\kappa_{M_{j}}f_{q_{1}}(x_{1}p_{M})f_{\bar{q}_{2}}(x_{2}p_{M}). (8)

κBj\kappa_{B_{j}} and κMj\kappa_{M_{j}} carry the information of pTp_{T}-independent combination dynamics. In order to determine their forms, we express momentum distributions of hadrons in another form

fBj(pB)\displaystyle f_{B_{j}}\left(p_{B}\right) =NBjfBj(n)(pB),\displaystyle=N_{B_{j}}\,f_{B_{j}}^{\left(n\right)}\left(p_{B}\right), (9)
fMj(pM)\displaystyle f_{M_{j}}\left(p_{M}\right) =NMjfMj(n)(pM).\displaystyle=N_{M_{j}}\,f_{M_{j}}^{\left(n\right)}\left(p_{M}\right). (10)

NBjN_{B_{j}} and NMjN_{M_{j}} are numbers of BjB_{j} and MjM_{j}, respectively. fBj(n)(pB)f_{B_{j}}^{\left(n\right)}\left(p_{B}\right) and fMj(n)(pM)f_{M_{j}}^{\left(n\right)}\left(p_{M}\right) are distribution functions normalized to one when integrating over momentum, which can be obtained by those of quarks and antiquarks,

fBj(n)(pB)\displaystyle f_{B_{j}}^{\left(n\right)}\left(p_{B}\right) =ABjfq1(n)(x1pB)fq2(n)(x2pB)fq3(n)(x3pB),\displaystyle=A_{B_{j}}\,f_{q_{1}}^{\left(n\right)}\left(x_{1}p_{B}\right)f_{q_{2}}^{\left(n\right)}\left(x_{2}p_{B}\right)f_{q_{3}}^{\left(n\right)}\left(x_{3}p_{B}\right), (11)
fMj(n)(pM)\displaystyle f_{M_{j}}^{\left(n\right)}\left(p_{M}\right) =AMjfq1(n)(x1pM)fq¯2(n)(x2pM).\displaystyle=A_{M_{j}}f_{q_{1}}^{\left(n\right)}\left(x_{1}p_{M}\right)f_{\bar{q}_{2}}^{\left(n\right)}\left(x_{2}p_{M}\right). (12)

Here, fqi(n)(p)=fqi(p)/Nqif_{q_{i}}^{\left(n\right)}\left(p\right)=f_{q_{i}}\left(p\right)/N_{q_{i}} is the normalized distribution function of quark qiq_{i}. Normalization coefficients ABjA_{B_{j}} and AMjA_{M_{j}} are defined as

ABj1\displaystyle A_{B_{j}}^{-1} =fq1(n)(x1pB)fq2(n)(x2pB)fq3(n)(x3pB)𝑑pB,\displaystyle=\int f_{q_{1}}^{\left(n\right)}\left(x_{1}p_{B}\right)f_{q_{2}}^{\left(n\right)}\left(x_{2}p_{B}\right)f_{q_{3}}^{\left(n\right)}\left(x_{3}p_{B}\right)dp_{B}, (13)
AMj1\displaystyle A_{M_{j}}^{-1} =fq1(n)(x1pM)fq¯2(n)(x2pM)𝑑pM.\displaystyle=\int f_{q_{1}}^{\left(n\right)}\left(x_{1}p_{M}\right)f_{\bar{q}_{2}}^{\left(n\right)}\left(x_{2}p_{M}\right)dp_{M}. (14)

Substituting Eqs. (11) and (12) into Eqs. (9) and (10) and then comparing them with Eqs. (7) and (8), we obtain

NBj\displaystyle N_{B_{j}} =Nq1Nq2Nq3κBjABj,\displaystyle=N_{q_{1}}N_{q_{2}}N_{q_{3}}\frac{\kappa_{B_{j}}}{A_{B_{j}}}, (15)
NMj\displaystyle N_{M_{j}} =Nq1Nq¯2κMjAMj.\displaystyle=N_{q_{1}}N_{\bar{q}_{2}}\frac{\kappa_{M_{j}}}{A_{M_{j}}}. (16)

From above two equations, we can read out the physical meaning of κBj\kappa_{B_{j}} and κMj\kappa_{M_{j}}. It is obvious that κBj/ABj\kappa_{B_{j}}/A_{B_{j}} denotes the momentum-integrated probability of q1q2q3q_{1}q_{2}q_{3} forming a baryon BjB_{j} and κMj/AMj\kappa_{M_{j}}/A_{M_{j}} denotes that of q1q¯2q_{1}\bar{q}_{2} forming a meson MjM_{j}. Two probabilities can be further decomposed as

κBjABj\displaystyle\frac{\kappa_{B_{j}}}{A_{B_{j}}} Pq1q2q3Bj=CBjNiterN¯BNq3,\displaystyle\equiv P_{q_{1}q_{2}q_{3}\rightarrow B_{j}}=C_{B_{j}}N_{iter}\frac{\overline{N}_{B}}{N_{q}^{3}}, (17)
κMjAMj\displaystyle\frac{\kappa_{M_{j}}}{A_{M_{j}}} Pq1q¯2Mj=CMjN¯MNqNq¯.\displaystyle\equiv P_{q_{1}\bar{q}_{2}\rightarrow M_{j}}=C_{M_{j}}\frac{\overline{N}_{M}}{N_{q}N_{\bar{q}}}. (18)

Taking meson for example, N¯M/NqNq¯\overline{N}_{M}/N_{q}N_{\bar{q}} is used to denote the averaged probability of a qq¯q\bar{q} pair forming a meson. Here, NqNq¯N_{q}N_{\bar{q}} is the all possible number of qq¯q\bar{q} pair, where Nq=Nu+Nd+NsN_{q}=N_{u}+N_{d}+N_{s} is the number of all quarks and Nq¯=Nu¯+Nd¯+Ns¯N_{\bar{q}}=N_{\bar{u}}+N_{\bar{d}}+N_{\bar{s}} is that of all antiquarks. N¯M\overline{N}_{M} is the average number of mesons produced by the hadronization of quark system with given quark number NqN_{q} and antiquark number Nq¯N_{\bar{q}}. The factor CMjC_{M_{j}} denotes further sophisticated tune for the production probability of MjM_{j} on the basis of the averaged probability N¯M/NqNq¯\overline{N}_{M}/N_{q}N_{\bar{q}}. The baryon formula is similar to meson except a factor NiterN_{iter}. This factor is to assure q1q2q3NiterNq1Nq2Nq3=Nq3\sum_{q_{1}q_{2}q_{3}}N_{iter}N_{q_{1}}N_{q_{2}}N_{q_{3}}=N_{q}^{3} and equals to 1,3,61,3,6 for q1q2q3q_{1}q_{2}q_{3} with three identical flavor, two identical flavor, and three different flavors, respectively. Finally, we obtain yield formulas of hadrons

NBj\displaystyle N_{B_{j}} =CBjNiterNq1Nq2Nq3N¯BNq3,\displaystyle=C_{B_{j}}N_{iter}N_{q_{1}}N_{q_{2}}N_{q_{3}}\frac{\overline{N}_{B}}{N_{q}^{3}}, (19)
NMj\displaystyle N_{M_{j}} =CMjNq1Nq¯2N¯MNqNq¯.\displaystyle=C_{M_{j}}N_{q_{1}}N_{\bar{q}_{2}}\frac{\overline{N}_{M}}{N_{q}N_{\bar{q}}}. (20)

N¯B\overline{N}_{B} and N¯M\overline{N}_{M} are functions of NqN_{q} and Nq¯N_{\bar{q}} Song and Shao (2013),

N¯M\displaystyle\overline{N}_{M} =x2[1z(1+z)a+(1z)a(1+z)a(1z)a],\displaystyle=\frac{x}{2}\left[1-z\frac{\left(1+z\right)^{a}+\left(1-z\right)^{a}}{\left(1+z\right)^{a}-\left(1-z\right)^{a}}\right], (21)
N¯B\displaystyle\overline{N}_{B} =xz3(1+z)a(1+z)a(1z)a,\displaystyle=\frac{xz}{3}\frac{\left(1+z\right)^{a}}{\left(1+z\right)^{a}-\left(1-z\right)^{a}}, (22)
N¯B¯\displaystyle\overline{N}_{\bar{B}} =xz3(1z)a(1+z)a(1z)a,\displaystyle=\frac{xz}{3}\frac{\left(1-z\right)^{a}}{\left(1+z\right)^{a}-\left(1-z\right)^{a}}, (23)

where x=Nq+Nq¯x=N_{q}+N_{\bar{q}} and z=(NqNq¯)/xz=\left(N_{q}-N_{\bar{q}}\right)/x. Parameter a=1+(N¯M/N¯B)z=0/3a=1+\left(\overline{N}_{M}/\overline{N}_{B}\right)_{z=0}/3 denotes the production competition of baryon to meson and is tuned to be a4.86±0.1a\approx 4.86\pm 0.1 according to our recent work Shao et al. (2017).

In this paper, we only consider the production of the ground state JP=0, 1J^{P}=0^{-},\,1^{-} mesons and JP=(1/2)+,(3/2)+J^{P}=(1/2)^{+},\,(3/2)^{+} baryons in flavor SU(3) group. In meson formation, we introduce a parameter RV/PR_{V/P} to describe the relative weight of a quark-antiquark pair forming the state of spin 1 to that of spin 0. Here, we take RV/P=0.55±0.05R_{V/P}=0.55\pm 0.05 in order to reproduce the measured K/KK^{*}/K and ϕ/K\phi/K data in high energy reactions Adam et al. (2016); Acharya et al. (2020). Factor CMjC_{M_{j}} is then parameterized as

CMj={11+RV/Pfor JP=0 mesonsRV/P1+RV/Pfor JP=1 mesons.C_{M_{j}}=\left\{\begin{array}[]{ll}\frac{1}{1+R_{V/P}}&\text{for }J^{P}=0^{-}\textrm{ mesons}\\ \frac{R_{V/P}}{1+R_{V/P}}&\textrm{for }J^{P}=1^{-}\textrm{ mesons}.\end{array}\right. (24)

In baryon formation, we introduce a parameter RD/OR_{D/O} to describe the relative weight of spin 3/23/2 state to 1/21/2 state for three quark combination. We take RD/O=0.5±0.04R_{D/O}=0.5\pm 0.04 by fitting the experimental data of Ξ/Ξ\Xi^{*}/\Xi and Σ/Λ\Sigma^{*}/\Lambda in high energy collisions Adamova et al. (2017). Then we have for CBjC_{B_{j}}

CBj={11+RD/Ofor JP=(1/2)+ baryonsRD/O1+RD/Ofor JP=(3/2)+ baryons,C_{B_{j}}=\left\{\begin{array}[]{ll}\frac{1}{1+R_{D/O}}&\textrm{for }J^{P}=({1}/{2})^{+}\textrm{ baryons}\\ \frac{R_{D/O}}{1+R_{D/O}}&\textrm{for }J^{P}=({3}/{2})^{+}\textrm{ baryons},\end{array}\right. (25)

except that CΛ=CΣ0=1/(2+RD/O),CΣ0=RD/O/(2+RD/O),CΔ++=CΔ=CΩ=1C_{\Lambda}=C_{\Sigma^{0}}={1}/{(2+R_{D/O})},\leavevmode\nobreak\ C_{\Sigma^{*0}}={R_{D/O}}/{(2+R_{D/O})},\leavevmode\nobreak\ C_{\Delta^{++}}=C_{\Delta^{-}}=C_{\Omega^{-}}=1.

Taking fqi(p)f_{q_{i}}\left(p\right) as model inputs, we can calculate fh(p)f_{h}\left(p\right) and NhN_{h} of hadrons directly produced at hadronization. Finally, we take the decay contribution of short-life resonances into account according to experimental measurements, and obtain results of final-state hadrons

fhj(final)(p)=fhj(p)+ij𝑑pfhi(p)Dij(p,p),f_{h_{j}}^{\left(final\right)}\left(p\right)=f_{h_{j}}\left(p\right)+\sum_{i\neq j}\int dp^{\prime}f_{h_{i}}\left(p^{\prime}\right)D_{ij}\left(p^{\prime},p\right), (26)

where the decay function Dij(p,p)D_{ij}\left(p^{\prime},p\right) is determined by the decay kinematics and decay branch ratios Olive et al. (2014).

III Strangeness neutralization in heavy-ion collisions

In relativistic heavy-ion collisions, strange quark and antiquark are always created in pair in collisions and therefore strangeness is globally conserved. However, for a finite kinetic region such as the mid-rapidity region, the strangeness neutralization is not so explicit, in particular, at low collision energies. In this section, using the precise data for yield densities and pTp_{T} spectra of identified hadrons Adler et al. (2004); Adams et al. (2007); Abelev et al. (2009a, b); Aggarwal et al. (2011); Adamczyk et al. (2017); Adam et al. (2020), we study the local strangeness in the midrapidity region in Au+Au collisions at STAR BES energies.

III.1 strangeness at mid-rapidity

In this subsection, we estimate strangeness density dNs¯/dydNs/dydN_{\bar{s}}/dy-dN_{s}/dy in the mid-rapidity region in relativistic heavy-ion collisions. We write Ns¯NsN_{\bar{s}}-N_{s} for short in the following. Since experimental measurements are mainly for hadrons in ground state in flavor SU(3), we estimate the strangeness by yield densities of the following hadrons

Ns¯Ns\displaystyle N_{\bar{s}}-N_{s}
=(K++K0+K++K0)\displaystyle=\left(K^{+}+K^{0}+K^{*+}+K^{*0}\right)
(K+K¯0+K+K¯0)\displaystyle-\left(K^{-}+\bar{K}^{0}+K^{*-}+\bar{K}^{*0}\right) (27)
(Λ+Σ0,±+Σ0,±+2Ξ,0+2Ξ,0+3Ω)\displaystyle-\left(\Lambda+\Sigma^{0,\pm}+\Sigma^{*0,*\pm}+2\Xi^{-,0}+2\Xi^{*-,0}+3\Omega^{-}\right)
+(Λ¯+Σ¯0,+Σ¯0,+2Ξ¯+,0+2Ξ¯+,0+3Ω¯+).\displaystyle+\left(\bar{\Lambda}+\bar{\Sigma}^{0,\mp}+\bar{\Sigma}^{*0,*\mp}+2\bar{\Xi}^{+,0}+2\bar{\Xi}^{*+,0}+3\bar{\Omega}^{+}\right).

For convenience, we use hh to denote dNh/dydN_{h}/dy. The contribution of baryons with different charge states is also abbreviated, e.g., Σ0,±Σ0+Σ+Σ+\Sigma^{0,\pm}\equiv\Sigma^{0}+\Sigma^{-}+\Sigma^{+}. We note that the contribution of higher mass resonances can be effectively included if we identify above ground state hadrons as the measured ones.

The strangeness in meson sector can be calculated as

K++K0+K++K0KK¯0KK¯0\displaystyle K^{+}+K^{0}+K^{*+}+K^{*0}-K^{-}-\bar{K}^{0}-K^{*-}-\bar{K}^{*0}
=(K+K)final+(K0K¯0)final,\displaystyle=\left(K^{+}-K^{-}\right)_{final}+\left(K^{0}-\bar{K}^{0}\right)_{final}, (28)

where we use the subscript finalfinal to denote that K+KK^{+}-K^{-} and K0K¯0K^{0}-\bar{K}^{0} have received the decay contribution of KK^{*}. Since neutral kaons are not measured, we use the approximation

(K+K)final(K0K¯0)final.\left(K^{+}-K^{-}\right)_{final}\approx\left(K^{0}-\bar{K}^{0}\right)_{final}. (29)

For strangeness contained in hyperons with one strange quark, we decompose them into two parts: experimentally measured net-Λ\Lambda and experimentally un-measured net-Σ±\Sigma^{\pm}. Using the property of S&EM decays for hyperons Olive et al. (2014), we have

(Λ+Σ0,±+Σ0,±)(Λ¯+Σ¯0,+Σ¯0,)\displaystyle\left(\Lambda+\Sigma^{0,\pm}+\Sigma^{*0,*\pm}\right)-\left(\bar{\Lambda}+\bar{\Sigma}^{0,\mp}+\bar{\Sigma}^{*0,*\mp}\right)
=(ΛΛ¯)final+(Σ±Σ¯)final\displaystyle=\left(\Lambda-\bar{\Lambda}\right)_{final}+\left(\Sigma^{\pm}-\bar{\Sigma}^{\mp}\right)_{final} (30)

with

(ΛΛ¯)final\displaystyle\left(\Lambda-\bar{\Lambda}\right)_{final} (31)
=(ΛΛ¯)+(Σ0Σ¯0)+0.94(Σ±Σ¯)\displaystyle=\left(\Lambda-\bar{\Lambda}\right)+\left(\Sigma^{0}-\bar{\Sigma}^{0}\right)+0.94\left(\Sigma^{*\pm}-\bar{\Sigma}^{*\mp}\right)
+0.88(Σ0Σ¯0)\displaystyle+0.88\left(\Sigma^{*0}-\bar{\Sigma}^{*0}\right)

and

(Σ±Σ¯)final\displaystyle\left(\Sigma^{\pm}-\bar{\Sigma}^{\mp}\right)_{final} (32)
=(Σ±Σ¯)+0.06(Σ±Σ¯)+0.12(Σ0Σ¯0).\displaystyle=\left(\Sigma^{\pm}-\bar{\Sigma}^{\mp}\right)+0.06\left(\Sigma^{*\pm}-\bar{\Sigma}^{*\mp}\right)+0.12\left(\Sigma^{*0}-\bar{\Sigma}^{*0}\right).

Applying our formula of hadronic yields in Sec. II, we obtain

(Σ±Σ¯)final(ΛΛ¯)final=1.1+0.68RD/O+0.099RD/O21.096+2.62RD/O+RD/O20.55.\frac{\left(\Sigma^{\pm}-\bar{\Sigma}^{\mp}\right)_{final}}{\left(\Lambda-\bar{\Lambda}\right)_{final}}=\frac{1.1+0.68R_{D/O}+0.099R_{D/O}^{2}}{1.096+2.62R_{D/O}+R_{D/O}^{2}}\approx 0.55. (33)

For strangeness contained in hyperons with two strange quarks, after considering S&EM decays, we have

(Ξ,0Ξ¯+,0)+(Ξ,0Ξ¯+,0)=(Ξ,0Ξ¯+,0)final\left(\Xi^{-,0}-\bar{\Xi}^{+,0}\right)+\left(\Xi^{*-,0}-\bar{\Xi}^{*+,0}\right)=\left(\Xi^{-,0}-\bar{\Xi}^{+,0}\right)_{final} (34)

and we use the approximation

(ΞΞ¯+)final(Ξ0Ξ¯0)final.\left(\Xi^{-}-\bar{\Xi}^{+}\right)_{final}\approx\left(\Xi^{0}-\bar{\Xi}^{0}\right)_{final}. (35)

By the sum over the strangeness in meson and baryon sectors, we obtain the net-strangeness of the system

Ns¯Ns\displaystyle N_{\bar{s}}-N_{s} 2(K+K)final1.57(ΛΛ¯)final\displaystyle\approx 2\left(K^{+}-K^{-}\right)_{final}-1.57\left(\Lambda-\bar{\Lambda}\right)_{final}
4(ΞΞ¯+)final3(ΩΩ¯+),\displaystyle-4\left(\Xi^{-}-\bar{\Xi}^{+}\right)_{final}-3\left(\Omega^{-}-\bar{\Omega}^{+}\right), (36)

where subscript finalfinal denotes the yield including S&EM decay contributions.

The total number of strange quarks and strange anti-quarks is

Ns¯+Ns\displaystyle N_{\bar{s}}+N_{s} 2(K++K)final+1.57(Λ+Λ¯)final\displaystyle\approx 2\left(K^{+}+K^{-}\right)_{final}+1.57\left(\Lambda+\bar{\Lambda}\right)_{final}
+4(Ξ+Ξ¯+)final+3(Ω+Ω¯+)\displaystyle+4\left(\Xi^{-}+\bar{\Xi}^{+}\right)_{final}+3\left(\Omega^{-}+\bar{\Omega}^{+}\right)
+2ϕ+2(23η+13η).\displaystyle+2\text{$\phi$}+2\left(\frac{2}{3}\eta+\frac{1}{3}\eta^{\prime}\right). (37)

Because the strangeness Ns¯NsN_{\bar{s}}-N_{s} is explicitly dependent on collision energies and collision centralities, we define the relative asymmetry factor

zS=Ns¯NsNs¯+Ns,z_{S}=\frac{N_{\bar{s}}-N_{s}}{N_{\bar{s}}+N_{s}}, (38)

which is convenient to compare results in different situations.

Table 1: Strangeness asymmetry factor zSz_{S} calculated by yield data of strange hadrons and anti-hadrons in central Au+Au collisions Adler et al. (2004); Adams et al. (2007); Abelev et al. (2009a, b); Aggarwal et al. (2011); Adamczyk et al. (2016b); Adam et al. (2020).
sNN\sqrt{s_{NN}} (GeV) K+,K^{+,-} Λ(Λ¯)\Lambda\left(\bar{\Lambda}\right) Ξ(Ξ¯+)\Xi^{-}\left(\bar{\Xi}^{+}\right) Ω(Ω¯+)\Omega^{-}\left(\bar{\Omega}^{+}\right) ϕ\phi zSz_{S}
200 48.9±6.345.7±5.2\begin{array}[]{c}48.9\pm 6.3\\ 45.7\pm 5.2\end{array} 16.7±1.112.7±0.9\begin{array}[]{c}16.7\pm 1.1\\ 12.7\pm 0.9\end{array} 2.17±0.21.83±0.2\begin{array}[]{c}2.17\pm 0.2\\ 1.83\pm 0.2\end{array} 0.53±0.060.53\pm 0.06 7.95±0.117.95\pm 0.11 0.004±0.06-0.004\pm 0.06
62.4 37.6±2.732.4±2.3\begin{array}[]{c}37.6\pm 2.7\\ 32.4\pm 2.3\end{array} 15.7±2.38.3±1.1\begin{array}[]{c}15.7\pm 2.3\\ 8.3\pm 1.1\end{array} 1.63±0.21.03±0.11\begin{array}[]{c}1.63\pm 0.2\\ 1.03\pm 0.11\end{array} 0.212±0.0280.167±0.027\begin{array}[]{c}0.212\pm 0.028\\ 0.167\pm 0.027\end{array} 3.52±0.083.52\pm 0.08 0.019±0.04-0.019\pm 0.04
39 32.0±2.925.0±2.3\begin{array}[]{c}32.0\pm 2.9\\ 25.0\pm 2.3\end{array} 11.02±0.033.82±0.01\begin{array}[]{c}11.02\pm 0.03\\ 3.82\pm 0.01\end{array} 1.54±0.010.78±0.01\begin{array}[]{c}1.54\pm 0.01\\ 0.78\pm 0.01\end{array} 0.191±0.0060.139±0.004\begin{array}[]{c}0.191\pm 0.006\\ 0.139\pm 0.004\end{array} 3.38±0.033.38\pm 0.03 0.002±0.05-0.002\pm 0.05
27 31.1±2.822.6±2.0\begin{array}[]{c}31.1\pm 2.8\\ 22.6\pm 2.0\end{array} 11.67±0.042.75±0.01\begin{array}[]{c}11.67\pm 0.04\\ 2.75\pm 0.01\end{array} 1.57±0.010.598±0.006\begin{array}[]{c}1.57\pm 0.01\\ 0.598\pm 0.006\end{array} 0.154±0.0080.0972±0.0049\begin{array}[]{c}0.154\pm 0.008\\ 0.0972\pm 0.0049\end{array} 3.01±0.043.01\pm 0.04 0.006±0.05-0.006\pm 0.05
19.6 29.6±2.918.8±1.9\begin{array}[]{c}29.6\pm 2.9\\ 18.8\pm 1.9\end{array} 12.58±0.041.858±0.009\begin{array}[]{c}12.58\pm 0.04\\ 1.858\pm 0.009\end{array} 1.62±0.020.421±0.005\begin{array}[]{c}1.62\pm 0.02\\ 0.421\pm 0.005\end{array} 0.155±0.010.0811±0.0048\begin{array}[]{c}0.155\pm 0.01\\ 0.0811\pm 0.0048\end{array} 2.57±0.042.57\pm 0.04 0.0002±0.05-0.0002\pm 0.05
11.5 25.0±2.512.3±1.2\begin{array}[]{c}25.0\pm 2.5\\ 12.3\pm 1.2\end{array} 14.17±0.080.659±0.009\begin{array}[]{c}14.17\pm 0.08\\ 0.659\pm 0.009\end{array} 1.35±0.020.169±0.004\begin{array}[]{c}1.35\pm 0.02\\ 0.169\pm 0.004\end{array} 0.082±0.0120.0356±0.0052\begin{array}[]{c}0.082\pm 0.012\\ 0.0356\pm 0.0052\end{array} 1.72±0.041.72\pm 0.04 0.004±0.05-0.004\pm 0.05
7.7 20.8±1.77.7±0.6\begin{array}[]{c}20.8\pm 1.7\\ 7.7\pm 0.6\end{array} 15.3±0.110.193±0.006\begin{array}[]{c}15.3\pm 0.11\\ 0.193\pm 0.006\end{array} 1.19±0.030.0667±0.0044\begin{array}[]{c}1.19\pm 0.03\\ 0.0667\pm 0.0044\end{array} 0.0271±0.00480.0075±0.0013\begin{array}[]{c}0.0271\pm 0.0048\\ 0.0075\pm 0.0013\end{array} 1.21±0.06\begin{array}[]{c}1.21\pm 0.06\end{array} 0.021±0.04-0.021\pm 0.04

In Table 1, we show results for zSz_{S} in most central Au+Au collisions at different collisions energies 111In calculations, we use the approximation of isospin symmetry between up and down quarks in Eqs. (29) and (35). If we consider the small asymmetry in number between up quarks and down quarks coming from colliding nucleons, we should modify Eq. (29) by multiplying a factor Nu/NdN_{u}/N_{d} and Eq. (35) by multiplying a factor (Nu/Nd)1(N_{u}/N_{d})^{-1} in right hand side of equation and the corresponding coefficients in Eqs. (36) and (37). According to the number of newborn quarks extracted in next section and net-quarks from participant nucleons, we obtain, for example, Nu/Nd0.99N_{u}/N_{d}\approx 0.99 at sNN=\sqrt{s_{NN}}= 200 GeV and Nu/Nd0.93N_{u}/N_{d}\approx 0.93 at 7.7 GeV. The resulting zSz_{S} are (-0.004, -0.018, -0.0014, -0.005, 0.002, 0.003, -0.009) at sNN=\sqrt{s_{NN}}= (200, 62.4, 39, 27, 19.6, 11.5, 7.7) GeV, respectively. They are very close to those shown in Tab. 1. Experimental data for yield densities of hadrons that are used in calculation are also presented Adler et al. (2004); Adams et al. (2007); Abelev et al. (2009a) Abelev et al. (2009b); Aggarwal et al. (2011); Abelev et al. (2009a) Adamczyk et al. (2016b); Adam et al. (2020). Because some data for Ω\Omega^{-} and ϕ\phi are results in 0-10% centrality, we re-scale them by multiplying a factor Npart(05%)/Npart(010%)N_{part}^{(0-5\%)}/N_{part}^{(0-10\%)} according to the participant nucleon number NpartN_{part}. We do the similar re-scaling for data of Ω\Omega^{-} in 0-60% centrality at sNN=\sqrt{s_{NN}}= 7.7 GeV. Because data of η\eta and η\eta^{\prime} mesons are usually absent, we neglect them and therefore the calculated zSz_{S} is over-estimated. We see that zSz_{S} at seven collision energies are quite small. Therefore, strangeness neutralization Ns=Ns¯N_{s}=N_{\bar{s}} is well satisfied in mid-rapidity region in heavy-ion collisions at RHIC energies.

III.2 pTp_{T} spectrum symmetry

In this subsection, we study the symmetry property between pTp_{T} spectrum of strange quark fs(pT)f_{s}\left(p_{T}\right) and that of strange antiquark fs¯(pT)f_{\bar{s}}\left(p_{T}\right). For this purpose, we choose Ω\Omega^{-} (Ω¯+\bar{\Omega}^{+}) which consists of only strange quarks (antiquarks). Using Eq. (7), we have

fΩ(3pT)\displaystyle f_{\Omega}\left(3p_{T}\right) =κΩfs3(pT),\displaystyle=\kappa_{\Omega}f_{s}^{3}\left(p_{T}\right), (39)
fΩ¯(3pT)\displaystyle f_{\bar{\Omega}}\left(3p_{T}\right) =κΩ¯fs¯3(pT)\displaystyle=\kappa_{\bar{\Omega}}f_{\bar{s}}^{3}\left(p_{T}\right) (40)

from which we have

fs¯(pT)fs(pT)=κΩ¯,Ω[fΩ¯(3pT)fΩ(3pT)]1/3[fΩ¯(3pT)fΩ(3pT)]1/3,\frac{f_{\bar{s}}\left(p_{T}\right)}{f_{s}\left(p_{T}\right)}=\kappa_{\bar{\Omega},\Omega}\left[\frac{f_{\bar{\Omega}}\left(3p_{T}\right)}{f_{\Omega}\left(3p_{T}\right)}\right]^{1/3}\propto\left[\frac{f_{\bar{\Omega}}\left(3p_{T}\right)}{f_{\Omega}\left(3p_{T}\right)}\right]^{1/3}, (41)

where κΩ¯,Ω=(κΩ¯/κΩ)1/3\kappa_{\bar{\Omega},\Omega}=\left(\kappa_{\bar{\Omega}}/\kappa_{\Omega}\right)^{1/3} is independent of pTp_{T} but is dependent on quark numbers. We emphasize that κΩ¯,Ω\kappa_{\bar{\Omega},\Omega} is not equal to one at nonzero net quark number.

Refer to caption
Figure 1: Ratio fs¯(pT)/fs(pT)f_{\bar{s}}\left(p_{T}\right)/f_{s}\left(p_{T}\right) in Au+Au collisions at different collision energies obtained from experimental data of Ω\Omega^{-} and Ω¯+\bar{\Omega}^{+} in central and semi-central collisions Adamczyk et al. (2016b); Adam et al. (2020) by Eq. (41).

Using data of pTp_{T} spectra for Ω\Omega^{-} and Ω¯+\bar{\Omega}^{+} in central Au+Au collisions Adamczyk et al. (2016b); Adam et al. (2020), we calculate the ratio fs¯(pT)/fs(pT)f_{\bar{s}}\left(p_{T}\right)/f_{s}\left(p_{T}\right) at different collision energies and present results in Fig. 1. Since we have shown Ns=Ns¯N_{s}=N_{\bar{s}} in the previous subsection, we multiply a proper constant before data of fΩ¯1/3(3pT)/fΩ1/3(3pT)f_{\bar{\Omega}}^{1/3}\left(3p_{T}\right)/f_{\Omega}^{1/3}\left(3p_{T}\right) to satisfy Ns=Ns¯N_{s}=N_{\bar{s}} and therefore we can directly compare fs¯(pT)/fs(pT)f_{\bar{s}}\left(p_{T}\right)/f_{s}\left(p_{T}\right) in/at different collision centralities/energies. Because of finite statistics of Ω\Omega^{-} and Ω¯+\bar{\Omega}^{+}, data points of fs¯(pT)/fs(pT)f_{\bar{s}}\left(p_{T}\right)/f_{s}\left(p_{T}\right) show a certain fluctuations around one. On the whole, we can see that fs¯(pT)=fs(pT)f_{\bar{s}}\left(p_{T}\right)=f_{s}\left(p_{T}\right) is a good approximation at the studied collision energies.

IV pTp_{T} spectra of hadrons

In this section, we use the quark combination model (QCM) in Sec. II to study pTp_{T} spectra of light-flavor hadrons in Au+Au collisions at RHIC energies. The inputs of model are pTp_{T} spectra of quarks and antiquarks at hadronization. Here, we take fs¯(pT)=fs(pT)f_{\bar{s}}\left(p_{T}\right)=f_{s}\left(p_{T}\right) based on the study of strangeness neutralization in Sec. III. We take fu¯(pT)=fd¯(pT)f_{\bar{u}}\left(p_{T}\right)=f_{\bar{d}}\left(p_{T}\right) for the newborn up and down anti-quarks. Because a part of up and down quarks comes from the colliding nucleons, pTp_{T} spectrum of up quarks is not exactly the same as that of down quarks. As discussed in [47], the relative difference in number between up and down quarks is only a few percentages. We have checked that pTp_{T} spectra of hadrons and yield ratios of anti-hadron to hadron studied in this paper are not sensitive to this small asymmetry. Therefore, in this paper, we take the approximation fu(pT)fd(pT)f_{u}\left(p_{T}\right)\approx f_{d}\left(p_{T}\right) in the mid-rapidity region.

In Table 2, we list all inputs and parameters of the model which are needed to calculate pTp_{T} spectra of hadrons. As introduced in model description in Sec. II, two parameters RV/PR_{V/P} and RD/OR_{D/O} are taken as 0.55 and 0.5, respectively, at all the studied collision energies. For three inputs, fs(pT)f_{s}\left(p_{T}\right) is fixed by using our model to fit experimental data of ϕ\phi, and fu(pT)f_{u}\left(p_{T}\right) (fu¯(pT)f_{\bar{u}}\left(p_{T}\right)) is fixed by experimental data of (anti-)baryons containing u(u¯)u(\bar{u}) such as (anti-)proton Abelev et al. (2009b); Adamczyk et al. (2017, 2016b), respectively. The extracted results for quark pTp_{T} spectra in Au+Au collisions for 0-5% centrality at six RHIC energies are shown in Fig. 2.

Table 2: Inputs and parameters of the model to calculation pTp_{T} spectra of hadrons.
input parameter
fu(pT)f_{u}(p_{T}) fu¯(pT)f_{\bar{u}}(p_{T}) fs(pT)f_{s}(p_{T}) RV/PR_{V/P} RO/DR_{O/D}
Refer to caption
Figure 2: pTp_{T} spectra of quarks at hadronization at mid-rapidity in Au+Au collisions for 0-5% centrality. Spectra at sNN=\sqrt{s_{NN}}= 7.7-27 GeV are scaled for clarity as shown in the figure.
Refer to caption
Figure 3: pTp_{T} spectra of hadrons at mid-rapidity in central Au+Au collisions at sNN=\sqrt{s_{NN}}= 200 GeV. Symbols are experimental data Abelev et al. (2006); Adams et al. (2007); Abelev et al. (2009a) and lines are results of our model. Spectra of some hadrons are scaled for clarity as shown in the figure.
Refer to caption
Figure 4: Same as Fig. 3 but for central Au+Au collisions at sNN=39\sqrt{s_{NN}}=39 GeV. Experimental data are from Adamczyk et al. (2016b, 2017); Adam et al. (2020).
Refer to caption
Figure 5: Same as Fig. 3 but for central Au+Au collisions at sNN=27\sqrt{s_{NN}}=27 GeV. Experimental data are from Adamczyk et al. (2016b, 2017); Adam et al. (2020).
Refer to caption
Figure 6: Same as Fig. 3 but for central Au+Au collisions at sNN=19.6\sqrt{s_{NN}}=19.6 GeV. Experimental data are from Adamczyk et al. (2016b, 2017); Adam et al. (2020).
Refer to caption
Figure 7: Same as Fig. 3 but for central Au+Au collisions at sNN=11.5\sqrt{s_{NN}}=11.5 GeV. Experimental data are from Adamczyk et al. (2016b, 2017); Adam et al. (2020).
Refer to caption
Figure 8: Same as Fig. 3 but for central Au+Au collisions at sNN=\sqrt{s_{NN}}= 7.7 GeV. Experimental data are from Adamczyk et al. (2016b, 2017); Adam et al. (2020).

In Fig. 3, we show the calculated results for pTp_{T} spectra of hadrons in central Au+Au collisions at sNN=\sqrt{s_{NN}}= 200 GeV and compare them with experimental data Abelev et al. (2006); Adams et al. (2007); Abelev et al. (2009a). We see that the agreement between our model results and experimental data is satisfactory. Although there exists many successful explanations on these experimental data in the framework of quark combination mechanism in previous works Fries et al. (2003); Greco et al. (2003); Hwa and Yang (2003); Chen and Ko (2006); Shao et al. (2005, 2009), here we would like to emphasize that the current EVC model can systematically explain these data in a simple way. Furthermore, the good agreement at top RHIC energy provides important basis for the application of our model to lower RHIC energies.

In Figs. 4, 5, 6, 7 and 8, we show results for pTp_{T} spectra of hadrons in central Au+Au collisions at sNN=\sqrt{s_{NN}}= 39, 27, 19.6, 11.5 and 7.7 GeV and compare with experimental data Adamczyk et al. (2016b, 2017); Adam et al. (2020). At sNN=\sqrt{s_{NN}}= 39, 27, 19.6, 11.5 GeV, we see a good agreement between our model results and experimental data Adamczyk et al. (2016b, 2017); Adam et al. (2020). In particular, we see that experimental data for baryons (pp, Λ\Lambda, Ξ\Xi, Ω\Omega) and meson ϕ\phi can be explained by the model very well. At sNN=\sqrt{s_{NN}}= 7.7 GeV, we also see that model results are in good agreement with available experimental data. However, compared with data in Figs. 3-7, the available data at sNN=\sqrt{s_{NN}}= 7.7 GeV cover smaller pTp_{T} range and Ω\Omega data in the most central collisions are absent. Therefore the comparison at sNN=\sqrt{s_{NN}}= 7.7 GeV is not as conclusive as those at other five higher energies. We should study this energy point further in the future when more precise data are available.

Using these systematic calculations and comparisons, we would like to emphasize the equal-velocity combination of quarks and antiquarks as the effective description at hadronization. This is manifested by the following two points. First, from panel (d) in Figs. 3-7, we see that experimental data of Ω\Omega and ϕ\phi can be perfectly explained by the same fs(pT)f_{s}\left(p_{T}\right) in a very simple way,

fΩ(pT)\displaystyle f_{\Omega}\left(p_{T}\right) =κΩfs3(pT/3),\displaystyle=\kappa_{\Omega}f_{s}^{3}\left(p_{T}/3\right), (42)
fϕ(pT)\displaystyle f_{\phi}\left(p_{T}\right) =κϕfs2(pT/2).\displaystyle=\kappa_{\phi}f_{s}^{2}\left(p_{T}/2\right). (43)

We call this property as the quark number scaling for hadronic pTp_{T} spectra. We have shown that this property not only exists in relativistic heavy-ion collisions but also exists in high-multiplicity pppp and pp-Pb collisions at LHC energies Song et al. (2017); Zhang et al. (2020); Song et al. (2020). Second, we see that data of Λ\Lambda and Ξ\Xi^{-} can be simultaneously explained by fs(pT)f_{s}\left(p_{T}\right) from ϕ\phi and fu(pT)f_{u}\left(p_{T}\right) from proton,

fΛ(pT)\displaystyle f_{\Lambda}\left(p_{T}\right) =κΛfu2(12+rpT)fs(r2+rpT),\displaystyle=\kappa_{\Lambda}f_{u}^{2}\left(\frac{1}{2+r}p_{T}\right)f_{s}\left(\frac{r}{2+r}p_{T}\right), (44)
fΞ(pT)\displaystyle f_{\Xi}\left(p_{T}\right) =κΞfu(11+2rpT)fs2(r1+2rpT),\displaystyle=\kappa_{\Xi}f_{u}\left(\frac{1}{1+2r}p_{T}\right)f_{s}^{2}\left(\frac{r}{1+2r}p_{T}\right), (45)

after further including the decay contribution of heavier baryons. Here, r=ms/mur=m_{s}/m_{u}. This is a clear support for the equal-velocity combination for quarks with different flavors.

V Hadronic yields and multi-particle correlations

In this section, we study the pTp_{T}-integrated yields of identified hadrons. In heavy-ion collisions at RHIC energies, the net baryon numbers deposited in the mid-rapidity region will influence the production of hadrons and antihadrons to a certain extent. One of the consequences for non-zero baryon number density is the asymmetry in yield between hadrons and their anti-particles. We study this yield asymmetry with our model by focusing on multi-particle yield correlations.

In Fig. 9, we show yield densities of hadrons and anti-hadrons 222Model results for yield densities of kaons are presented here. Because the direct combination u+s¯Ku+\bar{s}\to K in the current EVC model has energy conservation issue, we have to introduce further treatment to reconcile this such as we did in Ref. Gou et al. (2017). However, the strange quantum number conservation ensures that the number of the formed kaon can be correctly calculated in the current model. divided by participant nucleon number at mid-rapidity in central Au+Au collisions at different collision energies. Open symbols are experimental data Adams et al. (2004, 2007); Abelev et al. (2009b); Aggarwal et al. (2011); Adamczyk et al. (2017); Adam et al. (2020); Adcox et al. (2002b) and solid symbols are model results. Experimental data show that the yield split between K+K^{+} and KK^{-} in (a) is much smaller than that between pp and p¯\bar{p} in (b). Yield split between Λ\Lambda and Λ¯\bar{\Lambda} in (c) and that between Ξ\Xi^{-} and Ξ¯+\bar{\Xi}^{+} in (d) are larger than kaon split in (a) but are smaller than proton split in (b). Comparing with experimental data, we see that the model provides a globally good description for yield densities of hadrons and anti-hadrons.

Refer to caption
Figure 9: Hadronic yield densities divided by participant nucleon number at mid-rapidity in Au+Au collisions at different collision energies. Open symbols are experimental data Adams et al. (2004, 2007); Abelev et al. (2009b); Aggarwal et al. (2011); Adamczyk et al. (2017); Adam et al. (2020); Adcox et al. (2002b) and solid symbols are model results.

In order to further study the split in yield between hadrons and anti-hadrons, we calculate the ratio of yield for anti-hadron to that for hadron. Yield ratio is not sensitive to the numbers of quarks and antiquarks at hadronization and is also not sensitive to some hadronization details such as parameters RV/PR_{V/P} and RD/OR_{D/O} introduced in our model. Therefore, it can be used to provide a more direct test for the quark combination mechanism at hadronization in relativistic heavy-ion collisions. Experimentally measured protons and anti-protons contain complex decay contributions of heavier baryons and anti-baryons,

p(final)\displaystyle p^{(final)} =p+Δ+++23Δ++13Δ0+0.64Λ+0.517Σ+\displaystyle=p+\Delta^{++}+\frac{2}{3}\Delta^{+}+\frac{1}{3}\Delta^{0}+0.64\Lambda+0.517\Sigma^{+}
+0.64Σ0+0.633Σ++0.594Σ0+0.602Σ\displaystyle+0.64\Sigma^{0}+0.633\Sigma^{*+}+0.594\Sigma^{*0}+0.602\Sigma^{*-}
+0.64(Ξ0+Ξ+Ξ0+Ξ)+0.64Ω,\displaystyle+0.64\left(\Xi^{0}+\Xi^{-}+\Xi^{*0}+\Xi^{*-}\right)+0.64\Omega^{-}, (46)

where we use the particle name with superscript (final)\left(final\right) to denote the yield density of final-state hadron receiving the decay contributions and the particle name without superscript in the right hand side of the equation to denote yields of directly-produced hadrons by hadronization. The ratio p¯/p\bar{p}/p finally behaves as

p¯(final)p(final)(1z1+z)0.99a,\frac{\bar{p}^{\left(final\right)}}{p^{(final)}}\approx\left(\frac{1-z}{1+z}\right)^{0.99a}, (47)

where zz is net-quark fraction and the factor a4.86a\approx 4.86 is related to the baryon-meson production competition Song and Shao (2013); Shao et al. (2017), see Eq. (23) and texts below. For yields of kaons, we take the decay contributions of K(892)K^{*}\left(892\right) and ϕ\phi into account and have

K(final)K+(final)\displaystyle\frac{K^{-\left(final\right)}}{K^{+\left(final\right)}} =K+13K+23K0+0.489ϕK++13K++23K0+0.489ϕ\displaystyle=\frac{K^{-}+\frac{1}{3}K^{*-}+\frac{2}{3}K^{*0}+0.489\phi}{K^{+}+\frac{1}{3}K^{*+}+\frac{2}{3}K^{*0}+0.489\phi}
=1z1+z1+0.489Cϕλs1+λs(z1+z+0.489Cϕ1z1+z),\displaystyle=\frac{1-z}{1+z}\frac{1+0.489C_{\phi}\lambda_{s}}{1+\lambda_{s}\left(\frac{z}{1+z}+0.489C_{\phi}\frac{1-z}{1+z}\right)}, (48)

where λs=Ns/Nu¯\lambda_{s}=N_{s}/N_{\bar{u}}.

For yields of Λ\Lambda, Ξ\Xi^{-} and their anti-particles, we consider the S&EM decay contributions,

Λ¯(final)Λ(final)\displaystyle\frac{\bar{\Lambda}^{\left(final\right)}}{\Lambda^{\left(final\right)}} =Λ¯+Σ¯0+0.94(Σ¯+Σ¯+)+0.88Σ¯0Λ+Σ0+0.94(Σ++Σ)+0.88Σ0\displaystyle=\frac{\bar{\Lambda}+\bar{\Sigma}^{0}+0.94\left(\bar{\Sigma}^{*-}+\bar{\Sigma}^{*+}\right)+0.88\bar{\Sigma}^{*0}}{\Lambda+\Sigma^{0}+0.94\left(\Sigma^{*+}+\Sigma^{*-}\right)+0.88\Sigma^{*0}}
=(1z1+z)a1(1+λsz1+z)2,\displaystyle=\left(\frac{1-z}{1+z}\right)^{a-1}\left(1+\lambda_{s}\frac{z}{1+z}\right)^{-2}, (49)

and

Ξ¯+(final)Ξ(final)\displaystyle\frac{\bar{\Xi}^{+\left(final\right)}}{\Xi^{-\left(final\right)}} =Ξ¯++13Ξ¯++23Ξ¯0Ξ+13Ξ+23Ξ0\displaystyle=\frac{\bar{\Xi}^{+}+\frac{1}{3}\bar{\Xi}^{*+}+\frac{2}{3}\bar{\Xi}^{*0}}{\Xi^{-}+\frac{1}{3}\Xi^{*-}+\frac{2}{3}\Xi^{*0}}
=(1z1+z)a2(1+λsz1+z)1.\displaystyle=\left(\frac{1-z}{1+z}\right)^{a-2}\left(1+\lambda_{s}\frac{z}{1+z}\right)^{-1}. (50)

For Ω\Omega^{-}, we directly have

Ω¯+Ω=(1z1+z)a3.\frac{\bar{\Omega}^{+}}{\Omega^{-}}=\left(\frac{1-z}{1+z}\right)^{a-3}. (51)

In Eqs. (48-50), the power term [(1z)/(1+z)]α\left[\left(1-z\right)/\left(1+z\right)\right]^{\alpha} dominates the behavior of three ratios. Therefore, ratios K/K+K^{-}/K^{+}, p¯/p\bar{p}/p, Λ¯/Λ\bar{\Lambda}/\Lambda, Ξ¯+/Ξ\bar{\Xi}^{+}/\Xi^{-} and Ω¯+/Ω\bar{\Omega}^{+}/\Omega^{-} in our model are simply correlated with each other by the net-quark fraction zz.

Based on Eqs. (47-51), we can build several multi-hadron correlations as more sensitive tests of quark combination mechanism. In Fig. 10 (a), we firstly show the correlation between K/K+K^{-}/K^{+} and p¯/p\bar{p}/p. This correlation shows how the production of baryon and that of meson in heavy-ion collisions are simultaneously influenced by the baryon quantum number density characterized by net-quark fraction zz in our model. Symbols are experimental data at mid-rapidity in central and semi-central collisions Adcox et al. (2002b); Adams et al. (2004, 2007); Abelev et al. (2009b); Arsene et al. (2010); Aggarwal et al. (2011); Adamczyk et al. (2017); Adam et al. (2020). Error bars are the quadratic sum of statistical and systematic uncertainties. The dashed line is the result of QCM by Eqs. (47) and (48). Different from our previous work Song and Shao (2013) and previous experimental measurements Arsene et al. (2010), here we show the correlation in double-logarithmic coordinates in order to provide a full and clear presentation because ratio p¯/p\bar{p}/p changes much faster than K/K+K^{-}/K^{+}. We see that experimental data in double-logarithmic coordinates behave as almost a straight line and our model can well reproduce this correlation.

Refer to caption
Figure 10: Anti-hadron to hadron yield ratios in Au+Au collisions at different collision energies. Symbols are experimental data in central and semi-central (centrality <60%) collisions Adcox et al. (2002b); Adams et al. (2004, 2007); Abelev et al. (2009b); Arsene et al. (2010); Aggarwal et al. (2011); Adamczyk et al. (2017); Adam et al. (2020). Dashed lines are model results.

In Fig. 10 (b) we show the correlation between Λ¯/Λ\bar{\Lambda}/\Lambda and p¯/p\bar{p}/p. We see that experimental data, symbols in the figure, exhibit a linear behavior in double-logarithmic coordinates. This behavior can be perfectly reproduced in our model by Eqs. (47) and (49), see the dashed line. In addition, we note that ratio Λ¯/Λ\bar{\Lambda}/\Lambda decreases slower than p¯/p\bar{p}/p by a factor (1z)/(1+z)=Nq¯/Nq(1-z)/(1+z)=N_{\bar{q}}/N_{q}. This is because that, compared with p¯/p\bar{p}/p, ratio Λ¯/Λ\bar{\Lambda}/\Lambda involves a strangeness neutralization Ns¯/Ns=1N_{\bar{s}}/N_{s}=1

In Fig. 10 (c) we show the correlation between Ξ¯+/Ξ\bar{\Xi}^{+}/\Xi^{-} and p¯/p\bar{p}/p. Experimental data also exhibit a linear behavior in double-logarithmic coordinates. Since Ξ¯+/Ξ\bar{\Xi}^{+}/\Xi^{-} involves double effect of strangeness neutralization (Ns¯/Ns)2=1(N_{\bar{s}}/N_{s})^{2}=1, Ξ¯+/Ξ\bar{\Xi}^{+}/\Xi^{-} decreases slower than Λ¯/Λ\bar{\Lambda}/\Lambda as the function of p¯/p\bar{p}/p. Our model result Eqs. (47) and (50), the dashed line in the figure, can well describe data at sNN\sqrt{s_{NN}}\geq 11.5 GeV and slightly under-estimates Ξ¯+/Ξ\bar{\Xi}^{+}/\Xi^{-} at sNN=\sqrt{s_{NN}}= 7.7 GeV.

In Fig. 10 (d), we further show the correlation among mutl-strangeness hadrons Ω¯+/Ω\bar{\Omega}^{+}/\Omega^{-} and Ξ¯+/Ξ\bar{\Xi}^{+}/\Xi^{-}. Experimental data in double-logarithmic coordinates also exhibit a linear behavior. Because Ω\Omega^{-}( Ω¯+\bar{\Omega}^{+}) completely consists of strange (anti)quarks, ratio Ω¯+/Ω\bar{\Omega}^{+}/\Omega^{-} involves triple effect of strangeness neutralization and therefore it decreases slower than Ξ¯+/Ξ\bar{\Xi}^{+}/\Xi^{-}. The model result by Eqs. (50) and (51), the dashed line in the figure, can roughly describe data at sNN\sqrt{s_{NN}}\geq 11.5 GeV and under-estimates Ω¯+/Ω\bar{\Omega}^{+}/\Omega^{-} at sNN=\sqrt{s_{NN}}= 7.7 GeV to a certain extent.

Some discussions on above results are necessary. First, we emphasize the key physics in our model relating to multi-hadron correlations shown in Fig. 10. As shown by Eqs. (47)-(51), correlations among yield ratios of anti-hadron to hadron are not sensitive to absolute numbers of quarks and antiquarks at hadronization and non-perturbative parameters RV/PR_{V/P} and RD/OR_{D/O} introduced in the model. Therefore, these yield correlations are only related to two basic features of quark combination in our model. (1) free combination. Newborn quarks and antiquarks, net-quarks are all treated as individual (anti-)quarks and freely take part in combination. (2) flavor independent combination probability. We take N¯B/Nq3\overline{N}_{B}/N_{q}^{3} as the averaged probability of q1q2q3q_{1}q_{2}q_{3} forming a baryon and N¯M/(NqNq¯)\overline{N}_{M}/(N_{q}N_{\bar{q}}) as the averaged probability of q1q¯2q_{1}\bar{q}_{2} forming a meson. No sophisticated flavor correction is made at the moment. From Fig. 10, we see that such a global quark combination model provides a systematic description on multi-hadron yield correlations in Au+Au collisions, at least at sNN\sqrt{s_{NN}}\geq 11.5 GeV. Therefore, this is a clear signal of quark combination mechanism at hadronization in these collisions.

Second, the comparison between our model calculation and experimental data in Au+Au collisions at sNN=\sqrt{s_{NN}}= 7.7 GeV may indicate some physics beyond the key features of quark combination discussed above. As p¯/p\bar{p}/p and Λ¯/Λ\bar{\Lambda}/\Lambda ratios are reproduced, we see that model results for K/K+K^{-}/K^{+},Ξ¯+/Ξ\bar{\Xi}^{+}/\Xi^{-}and Ω¯+/Ω\bar{\Omega}^{+}/\Omega^{-} are smaller than experimental data to a certain extent. This may be related to the point (1) discussed above. In Au+Au collisions at low energy, the colliding nucleons may not break completely. A part of nucleon fragments may do not behave as the individual up/down quarks and freely take part in the combination with newborn quarks and antiquarks; Instead, they behave as diquarks and can form proton by combining with a up/down quark or form Λ\Lambda by combining with a strange quark. Because these net-quarks only contribute to proton and Λ\Lambda production, net-quark fraction zz used in Eqs. (48), (50) and (51) for K/K+K^{-}/K^{+}, Ξ¯+/Ξ\bar{\Xi}^{+}/\Xi^{-}and Ω¯+/Ω\bar{\Omega}^{+}/\Omega^{-} should be smaller than that in p¯/p\bar{p}/p and Λ¯/Λ\bar{\Lambda}/\Lambda. This consideration can increase ratios K/K+K^{-}/K^{+}, Ξ¯+/Ξ\bar{\Xi}^{+}/\Xi^{-}and Ω¯+/Ω\bar{\Omega}^{+}/\Omega^{-} at the given p¯/p\bar{p}/p and Λ¯/Λ\bar{\Lambda}/\Lambda ratios and therefore can qualitatively improve the description at sNN=\sqrt{s_{NN}}= 7.7 GeV in the current model. Such a sophisticated effect of net-quarks is worthwhile to be studied in detail in the future works.

VI properties of quark distributions

By studying experimental data of hadronic pTp_{T} spectra, we have obtained pTp_{T} spectra of quarks at hadronization, which are shown in Fig. 2. In this section, we study properties of these obtained quark distributions at hadronization.

We firstly calculate the ratio fs(pT)/fu¯(pT)f_{s}\left(p_{T}\right)/f_{\bar{u}}\left(p_{T}\right) and study its pTp_{T} dependence. Fig. 11 shows results in central Au+Au collisions at sNN=\sqrt{s_{NN}}= 200, 39, 27, 19.6 and 11.5 GeV. Result of fs(pT)/fu¯(pT)f_{s}\left(p_{T}\right)/f_{\bar{u}}\left(p_{T}\right) in pppp collisions at s=13\sqrt{s}=13 TeV Zhang et al. (2020) is also presented. We see that ratios in Au+Au collisions are obviously higher than that in pppp collisions. This means that the production of strange quarks in the studied pTp_{T} range in Au+Au collisions is significantly enhanced. In addition, we see that ratios in Au+Au collisions at these collision energies in the low pTp_{T} range (pT1.3p_{T}\lesssim 1.3 GeV/c) all increase with pTp_{T}. It is similar to pppp results. This property is related to the complex (non-)perturbative QCD evolution in connection with quark masses in partonic phase.

Refer to caption
Figure 11: Spectrum ratio fs(pT)/fu¯(pT)f_{s}\left(p_{T}\right)/f_{\bar{u}}\left(p_{T}\right) in central Au+Au collisions at sNN=\sqrt{s_{NN}}= 200, 39, 27, 19.6 and 11.5 GeV, and that in pppp collisions at s=13\sqrt{s}=13 TeV.

We note that the increase of the ratio at low pTp_{T} in heavy-ion collisions can be qualitatively understood by thermal statistics. In the case of thermal equilibrium such as the Boltzmann distribution dN/(pTdpT)exp(pT2+m2/T)dN/(p_{T}dp_{T})\propto\exp\left(-\sqrt{p_{T}^{2}+m^{2}}/T\right) in two-dimensional transverse momentum space, heavier mass will lead to flatter shape of the pTp_{T} distribution and thus lead to the increase of the ratio. However, Boltzmann distribution at hadronization temperature in the rest frame can not directly describe the obtained quark distributions in Fig. 2. We should further take into account the contribution of the collective radial flow in the prior partonic phase evolution in heavy-ion collisions. As an illustration, we consider a simple situation, that is, Boltzmann distribution in the two-dimensional transverse momentum space boosted with a radial flow velocity vv_{\perp}. In this case the distribution is

dNpTdpT1EE(v)exp[E(v)/T]\frac{dN}{p_{T}dp_{T}}\propto\frac{1}{E}E^{*}\left(v_{\perp}\right)\exp[-E^{*}\left(v_{\perp}\right)/T] (52)

with E(v)=γ(EvpT)E^{*}\left(v_{\perp}\right)=\gamma_{\perp}\left(E-v_{\perp}p_{T}\right) and E=pT2+m2E=\sqrt{p_{T}^{2}+m^{2}}. If we assume that quarks of different flavors at hadronization are thermalized and boosted with the same radial velocity, we can use the above formula to simultaneously describe the extracted fu¯(pT)f_{\bar{u}}\left(p_{T}\right) and fs(pT)f_{s}\left(p_{T}\right) in the low pTp_{T} range (pT1.3p_{T}\leq 1.3 GeV/c) in Fig. 2. According to our previous work Song et al. (2020), the hadronization temperature is taken as T=T= (0.164, 0.163, 0.162, 0.161, 0.156) GeV in central Au+Au collisions at sNN=\sqrt{s_{NN}}= (200, 39, 27, 19.6, 11.5) GeV, respectively. We obtain radial flow velocity v/cv_{\perp}/c\approx (0.39, 0.28, 0.27, 0.25, 0.23) at these collision energies.

We find that these results for radial flow velocity are consistent with our previous extraction by a hydrodynamics-motivated blast-wave model Schnedermann et al. (1993) fit of fs(pT)f_{s}\left(p_{T}\right) with the same hadronization temperature Song et al. (2020). We also find that fu¯(pT)f_{\bar{u}}\left(p_{T}\right) and fs(pT)f_{s}\left(p_{T}\right) extracted in this work can also be consistently described in blast-wave mode. Here, taking central Au+Au collisions at sNN=\sqrt{s_{NN}}= 19.6 GeV as an example, we show in Fig. 12 (a) the fit results of quark pTp_{T} spectra by Eq. (52) and those by blast-wave model at the same hadronization temperature 0.161 GeV and radial flow velocity 0.25 cc. We see that two fit methods give the consistent description on quark pTp_{T} spectra. In addition, we see from Fig. 12 (b) that the increase part of the ratio fs(pT)/fu¯(pT)f_{s}\left(p_{T}\right)/f_{\bar{u}}\left(p_{T}\right) in the low pTp_{T} range can be reasonably described.

Refer to caption
Figure 12: (a) Fit results for quark pTp_{T} spectra at hadronization in central Au+Au collisions at sNN=\sqrt{s_{NN}}= 19.6 GeV by Boltzmann formula Eq. (52) and by blast-wave model at the same hadronization temperature 0.161 GeV and radial flow velocity 0.25 cc. (b) ratio fs(pT)/fu¯(pT)f_{s}\left(p_{T}\right)/f_{\bar{u}}\left(p_{T}\right) by two fit methods. Symbols are quark pTp_{T} spectra and lines are fit results.

Fig. 11 also shows the ratio fs(pT)/fu¯(pT)f_{s}\left(p_{T}\right)/f_{\bar{u}}\left(p_{T}\right) globally increases with the decrease of collision energies. To study this energy dependence of strangeness, we calculate the strangeness factor

λs=NsNu¯\lambda_{s}=\frac{N_{s}}{N_{\bar{u}}} (53)

and present results in Fig. 13. Here, results of λs\lambda_{s} in central Au+Au collisions at sNN=\sqrt{s_{NN}}= 62.4 and 130 GeV are also shown. The uncertainty of λs\lambda_{s} is caused by that of experimental data for yield ratios such as K/πK/\pi, Λ/p\Lambda/p, Λ¯/p¯\bar{\Lambda}/\bar{p}. We note that these new results of λs\lambda_{s} are consistent with our previous works Shao et al. (2009); Sun et al. (2012); Shao et al. (2015). Compared with λs0.3\lambda_{s}\approx 0.3 in elementary collisions such as e+ee^{+}e^{-} and pppp reactions, we see that λs0.42\lambda_{s}\gtrsim 0.42 in heavy-ion collisions is obviously enhanced. We also see that λs\lambda_{s} increases as the decrease of collision energy.

Refer to caption
Figure 13: Strangeness factor λs=Ns/Nu¯\lambda_{s}=N_{s}/N_{\bar{u}} in central Au+Au collisions at different collision energies.

The dependence of λs\lambda_{s} on collision energy is related to the varied baryon quantum number density. In this paper, we study this energy dependence in the framework of thermodynamics for quark system. We consider a thermal system consisting of constituent quarks and antiquarks. In our quark combination model, constituent quarks and antiquarks are regarded as the effective degrees of freedom at hadronization and they freely combine into baryons and/or mesons at hadronization. Therefore, we can treat such a constituent quark system as the classical gas.

Because we study the production of hadrons in mid-rapidity range, we firstly discuss the case of grand-canonical ensemble. The number of quark under Fermi-Dirac statistics in the grand-canonical ensemble is

Nf\displaystyle N_{f} =gd3xd3p(2π)31exp[(Eμf)/T]+1\displaystyle=g\int\frac{d^{3}xd^{3}p}{\left(2\pi\right)^{3}}\frac{1}{\exp\left[\left(E-\mu_{f}\right)/T\right]+1}
=gVm2T2π2n=1(1)n+1nK2(nmT)exp(nμfT),\displaystyle=g\frac{Vm^{2}T}{2\pi^{2}}\sum_{n=1}\frac{\left(-1\right)^{n+1}}{n}K_{2}\left(n\frac{m}{T}\right)\exp\left(n\frac{\mu_{f}}{T}\right), (54)

where VV is system volume and TT is temperature. g=6g=6 is degeneracy factor of quark, mm is quark mass and μf\mu_{f} is chemical potential of quark ff. K2K_{2} is the modified Bessel function of the second kind. As discussions of local strangeness conservation in Sec. III, we set μs=μs¯=0\mu_{s}=\mu_{\bar{s}}=0 to get Ns=Ns¯N_{s}=N_{\bar{s}}. We assume the iso-spin symmetry between up and down quarks, then we have μu=μd=μu¯=μd¯=μB/3\mu_{u}=\mu_{d}=-\mu_{\bar{u}}=-\mu_{\bar{d}}=\mu_{B}/3. Then, the strangeness factor in grand-canonical ensemble (GCE) of free quark gas is

λs(GCE)\displaystyle\lambda_{s}^{\left(GCE\right)} =ms2n=1(1)n+1nK2(nmsT)mu2n=1(1)n+1nK2(nmuT)exp[nμB3T]\displaystyle=\frac{m_{s}^{2}\sum_{n=1}\frac{\left(-1\right)^{n+1}}{n}K_{2}\left(n\frac{m_{s}}{T}\right)}{m_{u}^{2}\sum_{n=1}\frac{\left(-1\right)^{n+1}}{n}K_{2}\left(n\frac{m_{u}}{T}\right)\exp[-n\frac{\mu_{B}}{3T}]}
λs,μB=0(GCE)exp[μB3T].\displaystyle\approx\lambda_{s,\mu_{B}=0}^{\left(GCE\right)}\exp[\frac{\mu_{B}}{3T}]. (55)

In the second line, we consider only the leading term n=1n=1, which is corresponding to the Boltzmann statistics. As we know, baryon chemical potential is increased as the decrease of the collision energy. Therefore, the collision energy dependence of λs\lambda_{s} is qualitatively understood.

For a demonstrative calculation for the collision energy dependence of λs(GCE)\lambda_{s}^{\left(GCE\right)}, we first apply Eq. (55) with the re-tuned mu=0.3m_{u}=0.3 GeV and ms=0.54m_{s}=0.54 GeV to give a proper strangeness at vanishing baryon chemical potential λs,μB=0(GCE)0.42\lambda_{s,\mu_{B}=0}^{\left(GCE\right)}\approx 0.42. Then, we apply Eq. (54) to fit the total quark number x=Nq+Nq¯x=N_{q}+N_{\bar{q}} and net-quark number xz=NqNq¯xz=N_{q}-N_{\bar{q}} integrated from Fig. 2 and obtain VV and μB\mu_{B} of quark system at hadronization. Here, the hadronization temperature is taken as before. Then we substitute μB\mu_{B} and TT into Eq. (55). The calculated results for λs(GCE)\lambda_{s}^{\left(GCE\right)} at RHIC energies are shown as triangles with the dashed line in Fig. 13. We see that the extracted λs\lambda_{s} in quark combination model can be explained by grand-canonical ensemble of quark gas system at sNN20\sqrt{s_{NN}}\gtrsim 20 GeV. At lower two collision energies, GCE results are higher than our extraction.

Considering the longitudinal rapidity space of heavy-ion collisions is finite, in particular, ybeam<2.5y_{beam}<2.5 at sNN11.5\sqrt{s_{NN}}\leq 11.5 GeV, the studied mid-rapidity range |y|<0.5|y|<0.5 is not a tiny fraction of the whole system, the hadron production in the midrapidity range may be influenced by effects of global charge conservation not only strangeness conservation but also baryon quantum number conservation. Therefore, grand-canonical treatment may be not perfectly suitable. We now consider the result of canonical ensemble for free quark system. We apply the method of canonical statistics Becattini and Ferroni (2004) to obtain the property of quark number distribution under the constraint of finite baryon quantum number and strangeness. We put the detailed derivation into Appendix A. The inputs of canonical ensemble are volume VV, temperature TT, and charges (BB,QQ,SS). The temperature is set to the hadronization temperature whose values at different collision energies are taken as before. As discussions in Sec. III, we take S=0S=0. VV, BB and QQ can be determined by fitting the quark and antiquark numbers integrated from Fig. 2. Results of λs(CE)\lambda_{s}^{\left(CE\right)} for canonical ensemble of quark system are shown as solid circles with the dotted line in Fig. 13. We see that λs(CE)\lambda_{s}^{\left(CE\right)} is also increased with the decrease of collision energy and is smaller than λs(GCE)\lambda_{s}^{\left(GCE\right)}, in particular, at low collision energies. Our extracted λs\lambda_{s} is roughly located in the middle of two ensembles.

VII Summary and discussions

In this paper, we have applied a quark combination model with equal-velocity combination approximation to systematically study the production of hadrons in Au+Au collisions at RHIC energies. The model applied in this work is motivated by our recent findings for the constituent quark number scaling property of hadronic pTp_{T} spectra in high energy pppp, ppA and AA collisions Song et al. (2017); Zhang et al. (2020); Song et al. (2020). After systematic study of pTp_{T} spectra and yields for hadrons, we found that our quark combination model provides a good explanation on the experimental data in Au+Au collisions at sNN11.5\sqrt{s_{NN}}\geq 11.5 GeV. This suggests that the constituent quark degrees of freedom still play important role even at low RHIC energy, which is closely related to the deconfinement at these collision energies.

By study of hadronic pTp_{T} spectra and yields in these collisions, we obtained pTp_{T} spectra and numbers of constituent quarks and antiquarks at hadronization. We calculated the net strangeness Ns¯NsN_{\bar{s}}-N_{s} at mid-rapidity and showed the strangeness neutralization is well satisfied at mid-rapidity in heavy-ion collisions at RHIC energies. We studied the spectrum ratio of strange quarks to newborn up/down quarks fs(pT)/fu¯(pT)f_{s}\left(p_{T}\right)/f_{\bar{u}}\left(p_{T}\right) and the pTp_{T} integrated number ratio λs=Ns/Nu¯\lambda_{s}=N_{s}/N_{\bar{u}} at different collision energies. We applied the basic thermal statistics for free constituent quark system to understand these properties of strange quarks relative to up/down quarks.

We emphasize the key physics in our model which are responsible for successfully explaining experimental data of hadronic pTp_{T} spectra and yields. First, the model takes the constituent quarks and antiquarks as the effective interface connecting the strongly-interacting system before hadronization and that after hadronization. We assume the constituent quarks and antiquarks as effective degrees of freedom for the strongly-interacting quark-gluon system just before hadronization. On the other hand, we take the constituent quark model to describe the static structure of hadrons in the ground state. In this scenario, the equal-velocity combination of these constituent quarks and antiquarks is a reasonable approximation and is indeed supported by the quark number scaling property for pTp_{T} spectra of hadrons observed in experiments Song et al. (2017); Zhang et al. (2020); Song et al. (2020). The study in this paper further showed that the equal-velocity combination can systematically describe the production of different kinds of hadrons in heavy-ion collisions at RHIC energies. Second, the model includes reasonable considerations for the unity of hadronization and the linear response property, see detailed discussions in Song and Shao (2013). This is very important to reproduce the multi-hadron yield correlations shown in Fig. 10. For example, in a naive inclusive view of combination, we have NΩNs3N_{\Omega}\propto N_{s}^{3} and NΩ¯Ns¯3N_{\bar{\Omega}}\propto N_{\bar{s}}^{3} and therefore Ω¯+/Ω(Ns¯/Ns)31\bar{\Omega}^{+}/\Omega^{-}\propto\left(N_{\bar{s}}/N_{s}\right)^{3}\approx 1 which is independent of collision energy. However, yield of Ω\Omega^{-} in our model is not only dependent on Ns3N_{s}^{3} but also dependent on the global system information shown as in Eq. (17); thus we have Ω¯+/Ω=[(1z)/(1+z)]a3\bar{\Omega}^{+}/\Omega^{-}=\left[(1-z)/(1+z)\right]^{a-3} in Eq. (51) which decreases with the decrease of collision energy.

In order to further test our EVC model, the following two aspects are deserved to study in depth in future works. The first is the effect of momentum correlations among quarks and antiquarks at hadronization. It is neglected in this paper, see Eqs. (5) and (6). In general, quarks and antiquarks at hadronization always have some momentum correlations. In heavy-ion collisions, the collective flow formed in parton phase will cause a certain correlation among momenta of quarks and antiquarks. This correlation not only influences the inclusive momentum spectra of hadrons to a certain extent but also influences multi-particle momentum correlations more directly. In the future work, we will study this effect by building sensitive physical observables in EVC model. The second is the production of short-lived resonances. In heavy-ion collisions, yield and momentum spectra of finally-observed resonances such as K(892)K^{*}(892) are strongly influenced by re-scatterings among hadrons. In the future work, we will systematically consider this hadronic rescattering effect and study the production mechanism of the short-lived resonances at hadronization.

VIII Acknowledgments

This work is supported in part by the National Natural Science Foundation of China under Grant No. 11975011, Shandong Province Natural Science Foundation under Grant Nos. ZR2019YQ06 and ZR2019MA053, and Higher Educational Youth Innovation Science and Technology Program of Shandong Province (2019KJJ010, 2020KJJ004).

Appendix A Quark number distribution in canonical ensemble

The probability of a single state in the canonical ensemble is

Pstate=1Z(𝑸)eEstateβδ𝑸,𝑸state,P_{state}=\frac{1}{Z\left(\bm{Q}\right)}e^{-E_{state}\beta}\delta_{\bm{Q},\bm{Q}_{state}}, (56)

where β=1/T\beta=1/T is the inverse temperature, EstateE_{state} is the energy of the state and 𝑸state\bm{Q}_{state} are abelian charges of the state

𝐐state\displaystyle\mathbf{Q}_{state} =\displaystyle= j=1K𝐪jNj.\displaystyle\sum_{j=1}^{K}\mathbf{q}_{j}N_{j}. (57)

Here NjN_{j} is the number of particle jj in the current state and 𝐪j=(Bj,Qj,Sj)\mathbf{q}_{j}=\left(B_{j},Q_{j},S_{j}\right) is the quantum number vector for the jj-th particle. KK is the number of particle species.

The multiplicity distribution of particles can be obtained from the generating function associated to the canonical partition function Z(𝑸)Z(\bm{Q}) Becattini and Ferroni (2004) and is expressed as

P({Nj})\displaystyle P\left(\left\{N_{j}\right\}\right) (58)
=1Z(𝑸)j=1K{{hnj}nj=1Nj[[zj(nj)]hnjnjhnjhnj!]}δ𝑸,j𝒒𝒋Nj,\displaystyle=\frac{1}{Z\left(\bm{Q}\right)}\,\prod_{j=1}^{K}\left\{\sum_{\left\{h_{n_{j}}\right\}}\prod_{n_{j}=1}^{N_{j}}\left[\frac{\left[z_{j}\left(n_{j}\right)\right]^{h_{n_{j}}}}{n_{j}^{h_{n_{j}}}h_{n_{j}}!}\right]\right\}\delta_{\bm{Q},\sum_{j}\bm{q_{j}}N_{j}},

where the summation takes different configurations for {hnj}\left\{h_{n_{j}}\right\} into account under the condition nj=1njhnj=Nj\sum_{n_{j}=1}^{\infty}n_{j}h_{n_{j}}=N_{j}.

zj(nj)\displaystyle z_{j}\left(n_{j}\right) =()nj+1gV(2π)3d3penβE\displaystyle=\left(\mp\right)^{n_{j}+1}\frac{gV}{\left(2\pi\right)^{3}}\int d^{3}p\,e^{-n\beta E}
=()nj+1gVm22π2njβK2(njβm).\displaystyle=\left(\mp\right)^{n_{j}+1}gV\frac{m^{2}}{2\pi^{2}n_{j}\beta}K_{2}(n_{j}\beta m). (59)

Now, we consider the thermal system consisting of constituent quarks and antiquarks. In our quark combination model, constituent quarks and antiquarks are regarded as the effective degrees of freedom at hadronization and they freely combine to form baryons and/or mesons at hadronization. Therefore, we can simply apply above formula to obtain the number distribution of these “free” constituent quarks and antiquarks under canonical statistics. Here, we consider up, down, strange quarks and their antiparticles. Index jj denotes u,d,su,d,s, u¯,d¯,s¯\bar{u},\bar{d},\bar{s} and K=6K=6.

Eq. (58) can be further denoted as

P({Nj})=1Z(𝑸)[j=1K𝒵(SNj)]δ𝑸,jNj𝒒𝒋,\displaystyle P\left(\left\{N_{j}\right\}\right)=\frac{1}{Z\left(\bm{Q}\right)}\left[\prod_{j=1}^{K}\mathcal{Z}\left(S_{N_{j}}\right)\right]\delta_{\bm{Q},\sum_{j}N_{j}\bm{q_{j}}}, (60)

where 𝒵(SNj)\mathcal{Z}\left(S_{N_{j}}\right) is cycle-index polynomial of symmetric group. 𝒵(SNj)\mathcal{Z}\left(S_{N_{j}}\right) can be numerically evaluated using the recurrence relation

𝒵(SN)=1Nl=1Nz(l)𝒵(SNl)\mathcal{Z}\left(S_{N}\right)=\frac{1}{N}\sum_{l=1}^{N}z\left(l\right)\mathcal{Z}\left(S_{N-l}\right) (61)

with 𝒵(S0)=1\mathcal{Z}\left(S_{0}\right)=1 and 𝒵(S1)=z(1)\mathcal{Z}\left(S_{1}\right)=z\left(1\right).

As discussed in Sec. III, we take strangeness neutralization S=0S=0 and therefore Ns=Ns¯N_{s}=N_{\bar{s}}. For constraints of baryon charge BB and electric charge QQ, we denote them as NuNu¯=B+QN_{u}-N_{\bar{u}}=B+Q and NdNd¯=2BQN_{d}-N_{\bar{d}}=2B-Q. Finally the joint distribution function of quark numbers and antiquark numbers is

P(Nd,Nu,Ns,Nd¯,Nu¯,Ns¯)\displaystyle P\left(N_{d},N_{u},N_{s},N_{\bar{d}},N_{\bar{u}},N_{\bar{s}}\right)
=1Z(𝑸)𝒵(SNd)𝒵(SNd¯)δNd,Nd¯+2BQ\displaystyle=\frac{1}{Z\left(\bm{Q}\right)}\mathcal{Z}\left(S_{N_{d}}\right)\mathcal{Z}\left(S_{N_{\bar{d}}}\right)\,\delta_{N_{d},N_{\bar{d}}+2B-Q}
×𝒵(SNu)𝒵(SNu¯)δNu,Nu¯+B+Q\displaystyle\,\,\,\,\,\times\mathcal{Z}\left(S_{N_{u}}\right)\mathcal{Z}\left(S_{N_{\bar{u}}}\right)\delta_{N_{u},N_{\bar{u}}+B+Q}
×𝒵(SNs)𝒵(SNs¯)δNs,Ns¯.\displaystyle\,\,\,\,\,\times\mathcal{Z}\left(S_{N_{s}}\right)\mathcal{Z}\left(S_{N_{\bar{s}}}\right)\delta_{N_{s},N_{\bar{s}}}. (62)

We obtain the averaged number of quarks by

N¯s\displaystyle\overline{N}_{s} ={Nqi}NsP(Nd,Nu,Ns,Nd¯,Nu¯,Ns¯),\displaystyle=\sum_{\left\{N_{q_{i}}\right\}}N_{s}P\left(N_{d},N_{u},N_{s},N_{\bar{d}},N_{\bar{u}},N_{\bar{s}}\right), (63)
N¯u¯\displaystyle\overline{N}_{\bar{u}} ={Nqi}Nu¯P(Nd,Nu,Ns,Nd¯,Nu¯,Ns¯),\displaystyle=\sum_{\left\{N_{q_{i}}\right\}}N_{\bar{u}}P\left(N_{d},N_{u},N_{s},N_{\bar{d}},N_{\bar{u}},N_{\bar{s}}\right), (64)

and calculate strangeness factor λs(CE)=N¯s/N¯u¯\lambda_{s}^{\left(CE\right)}=\overline{N}_{s}/\overline{N}_{\bar{u}} in canonical ensemble of free quark system.

References