logo

Quark mass scaling and properties of light-quark matter

Special Section on International Workshop on Nuclear Dynamics in Heavy-Ion Reactions (IWND2016)

Quark mass scaling and properties of light-quark matter

Zhen-Yan Lu
Guang-Xiong Peng
Shi-Peng Zhang
Marco Ruggieri
Vincenzo Greco
Nuclear Science and TechniquesVol.27, No.6Article number 148Published in print 20 Dec 2016Available online 31 Oct 2016
50500

We study the properties of two-flavor quark matter in the equivalent particle model. A new quark mass scaling at finite temperature is proposed and applied to the thermodynamics of two-flavor quark matter. It is found that the perturbative interaction has strong effect on quark matter properties at finite temperature and high density. The pressure at the minimum free energy per baryon is exactly zero. With increasing temperature, the energy per baryon increases while the free energy per baryon decreases.

Quark matterEquation of stateQuark mass scaling

1 Introduction

In many fields of nuclear physics, e.g. the Quantum Chromodynamics (QCD) phase diagram and structure of compact stars, one needs the equation of state of quark matter [1]. In principle, we have the fundamental theory of strong interactions, i.e. QCD. Due to the non-perturbative QCD interaction at relatively lower densities on one hand, and the so-called "sign problem" on the other hand, QCD can not be studied from first principles at finite baryon density. Therefore, one usually resort to various phenomenological models, e.g., Nambu and Jona-Lasinio model [2], perturbation model [3-5], field correlator method [6], quark-cluster model [7], and many other models [8-18].

A useful way to include interactions among quarks is to use medium-dependent quark masses. In one class of models, namely the quasiparticle models, the quark mass depends on chemical potential and/or temperature [19-27]. In another case, the quark mass depends on density and/or temperature [28-31]. These kinds of models were previously called the density-dependent (density- and temperature-dependent) quark mass models, but the thermodynamic treatment is inconsistent.

In the case of chemical-potential dependent masses, one can have thermodynamic consistency by adding an additional term to the thermodynamic potential density [20]. In the density-dependent case, it is now clear that the original chemical potentials should be replaced with effective ones when the quark masses become density- and temperature- dependent [32]. The most recent version is fully thermodynamically self-consistent, and is called the equivparticle model [33].

Another important issue in the equivparticle model is the quark mass scaling, i.e. how to parameterize the density dependence of the quark masses. Initially, people mainly emphasize quark confinement, and the interaction part in the quark mass is parameterized to be inversely proportional to the baryon number density [28] for light quarks, and later extended to strange quarks [29]. According to the in-medium chiral condensate, it was shown that the interaction quark mass should be inversely proportional to the cubic root of the density [14].

It is true that confinement interactions dominates at lower densities. With increasing densities, however, the perturbative interaction sets in, and becomes more and more important, and thus should be included. Recently, a new quark-mass scaling, considering both the confinement and perturbative interactions, was derived, and applied to the investigation of SQM and strange stars [33]. But the important temperature effect was not considered there.

The purpose of the present paper is to study the properties of two-flavor quark matter with the equivparticle model with the inclusion of both confinement and first-order perturbative interaction at finite temperature and density. In Sect. 2, we will discuss the quark mass scaling, while in Sect. 3, we give the necessary thermodynamic framework and numerical results. Finally we summarize the paper in Sect. 4.

2 Quark mass scaling

Originally, the light quark mass mq (q=u/d) was assumed to be inversely proportional to density, i.e. [28]

mu/d=B3nb, (1)

where B is the bag constant, and nb is the baryon number density of quark matter. It was soon extended to including strange quarks [30]

ms=ms0+B3nb. (2)

Based on the in-medium chiral condensates and linear confinement, a new scaling was derived [14]

mq=mq0+Dnb1/3. (3)

This cubic scaling has been extensively applied to the investigations of SQM-related physics, e.g. the QCD phase diagram [32], properties of SQM and slets [34] at zero and finite temperature [16, 35], the damping time scale due to the coupling of the viscosity and r mode [36], and the quark-diquark equation of state and compact star structure [37] etc. The energy per baryon of quark matter becomes infinite for small density, describing effectively quark confinement.

Although the confinement interaction is dominant at lower density, perturbative interactions become more and more important with increasing density. Recently, a new quark mass scaling considers both confinement and first-order perturbative interactions has been derived [33] as

mq=mq0+Dnb1/3+Cnb1/3. (4)

The cubic-root scaling in Eq. (3) was extended to include finite temperature [16] as

mq=mq0+Dnb1/3[18TλTcexp(λTcT)], (5)

where λ 1.6, and Tc is the critical temperature from where the confinement effect becomes ignorable.

The advantage of the extension to finite temperature in Eq. (5) is that it is in consistent with the string theory at lower temperature. An obvious disadvantage is that when one extends it to high temperature, it approaches to zero at T=Tc, which was previously regarded as a critical temperature. When T>Tc, the expression becomes negative, and one has to take a zero value for the interaction part. We can now write Eq. (5) in another form as

mq=mq0+Dnb1/3[1+8TλTcexp(λTcT)]1. (6)

The new expression in Eq. (6) equals to that in Eq. (5) at lower temperature. At higher temperature, however, they act very differently: the former becomes inversely proportional to temperature while the latter is zero.

Observing Eq. (4), we notice that, except for an adjustable constant, the confinement and perturbative interaction terms are reciprocal. Therefore, inspired by Eqs. (4) and (6) we assume that at finite temperature and density quark mass behaves:

mq=mq0+Dnb1/3(1+8TΛeΛ/T)1+Cnb1/3(1+8TΛeΛ/T), (7)

where Λ is a temperature scale parameter, comparable with ΛQCD. Comparing the low density limit of Eq. (7) with Eq. (6), we should have Λ=λTc. We thus take Λ=280 MeV in the present calculations.

The confinement parameter D and the perturbative strength C should be chosen to meet the condition that the minimum energy per baryon of light-quark matter is greater than 930 MeV at zero temperature, in order not to contradict with the conventional nuclear physics. According to our recent study, D1/2 should be bigger than 120 MeV [33] but smaller than 270 MeV [16], and C is smaller than 1 [33]. For the present case of the two-flavor quark matter, we take the modest values D1/2=160 MeV and C=0.6. This choice has the property that the confinement term decreases up to a flex temperature point at about 155 MeV.

In Fig. 1, we show the quark mass as a function of the temperature at the density nb=0.5 fm-3 for the parameters D1/2=160 MeV and C=0.6 with a solid curve. The confinement and perturbative contributions are also plotted, respectively, with dotted and dashed curves. We notice that with increasing temperature, the confinement term in the quark mass becomes less important than the perturbative one.

Fig. 1.
The temperature dependence of the quark mass at density nb=0.5 fm-3 with the parameters D1/2=160 MeV, C=0.6 and Λ=280 MeV.
pic

For the gluon mass, we, according to the first-order perturbative result, use

(mg/T)2=ηαθ(TTc), (8)

where Tc is the critical temperature of a pure SU(3) gluon gas. The step function, θ, is unity with a non-negative argument and zero otherwise. The parameter, η, is taken to be η=15 in the present calculations, while for the running coupling, we use a fast convergent expansion [42]:

α=β0β02ln(u/ΛT)+β1lnln(u/ΛT), (9)

where the beta coefficients are taken to be β0=11/2-Nf/3 and β1=51/4-(19/12)Nf. In order to include nonperturbative effects, the renormalization subtraction point is assumed to vary linearly with the temperature as

u/ΛT=c0+c1x (10)

with c0=1, c1=1/2, and x T/Tc.

To calculate the properties of quark matter in the equivparticle model, we need to use the density and temperature derivatives of the quark mass, which can be easily obtained from Eq. (7), giving

mqnb=D3nb4/3(1+8TΛeΛ/T)1+C3nb2/3(1+8TΛeΛ/T), (11)

and

mqT=8DΛnb1/3(1+ΛT)eΛ/T(1+8TΛeΛ/T)2+8Cnb1/3(1T+1Λ)exp(ΛT). (12)

The temperature derivative of the gluon mass can be similarly obtained from Eq. (8), i.e.,

dmgdT=mgT(1+x2dlnαdx), (13)

where

dlnαdx=c1αc0+c1x[β0+β1β01ln(c0+c1x)]. (14)

In literature, there are other forms of the quark mass scaling, e.g., an isospin term was considered in Ref. [38]; Eqs. (1) and (2) were extended to finite temperature by expansion to a Taylor series of temperature [39]; The one-gluon-exchange effect was considered [40]; The asymptotic freedom was explicitly considered by a Wood-Saxon factor [41]; etc. In the present paper, however, we use the quark mass formula in Eq. (7) to study the properties of light-quark matter.

3 Properties of two-flavor quark matter

For simplicity, we assume the system consisting of light quarks and gluons at finite temperature T. According to the equivalent particle model, the free particle contribution to the thermodynamic potential density can be written in the form

Ω0=Ω0++Ω0+Ω0g, (15)

where the equivalent quark (+) and antiquark (-) contributions are

Ω0±=dqT2π20ln[1+e(p2+mq2μ*)/T]p2dp, (16)

with the degeneracy factor dq=3(colors)× 2(spins)× 2(flavors)=12, while the contribution from gluons is

Ω0g=dgT2π20ln[1ep2+mg2/T]p2dp. (17)

The net quark baryon number density is then

nb=13(nq+nq), (18)

where nq+ and nq are, respectively, the quark/antiquark number densities:

nq±=dq2π20p2dpe(p2+mq2μ*)/T+1. (19)

We emphasize that the quark chemical potential becomes effective due to the density and temperature dependence of the quark mass. The effective chemical potential, μ*, is related to the actual chemical potential, μ, by

μ=μ*+13Ω0mqmqnb, (20)

where dmq/dnb is given in Eq. (11), and the derivative of Ω0, with respect to mq, is given by

Ω0mq=dqmq2π20[1e(p2+mq2μ*)/T+1+1e(p2+mq2+μ*)/T+1]p2dpp2+mq2. (21)

The pressure of light-quark matter at finite temperature is

P=Ω0+nbmqnbΩ0mq (22)

with the second term on the right hand side of the above equation arising due to the density dependence of the quark mass. The entropy density reads

S=Ω0Ti=q,gmiTΩ0mi, (23)

while the free energy density is given by

F=Ω0+3nbμ*. (24)

Therefore, the energy density can be obtained by substituting Eqs. (23) and (24) into the thermodynamic relation E=F+TS, which leads to

E=Ω0+3nbμ*+TS. (25)

For a given pair of the density, nb, and temperature, T, we can first solve Eq. (18) to obtain the effective chemical potential, μ*, and the thermodynamic properties can be calculated by Eqs. (22)–(25). Please note, in these expressions, the contribution from gluons has already been included.

In Fig. 2, we show the energy per baryon as functions of the density at the fixed temperature values of T=100 MeV, 150 MeV, and 200 MeV for the parameters D1/2=160 MeV and C=0.6. At lower density, the energy per baryon becomes infinitely large, indicating quark confinement. At large baryon density, the energy per baryon also becomes large because the perturbative interactions become gradually important.

Fig. 2.
The density dependence of the energy per baryon at different temperature values.
pic

The temperature dependence of the energy (dashed) and free energy (solid) per baryon are given in Fig. 3. We notice that the energy per baryon increases with temperature. However, the free energy per baryon decreases with increasing temperature, and eventually becomes negative. This is understandable with a view to the relation F=E-TS and the fact that the entropy term -TS dominates at large temperature.

Fig. 3.
The temperature dependence of the energy (dashed) and free energy (solid) per baryon at different densities.
pic

To check the thermodynamic consistency and demonstrate the effect of the perturbative interaction, we separately plot the density dependence of the energy (dashed curves) and free energy (solid) per baryon as a function of the density at the fixed temperature T=30 MeV for C=0 (open square) and C=0.6 (full square) in Fig. 4 where the points marked with a triangle is the minimum while the points marked with an open circle are the zero pressure. In Fig. 5, the pressure and the corresponding free energy per baryon at T=50 MeV are simultaneously shown. It is clearly seen that the pressure is negative when the density is lower than the density where the free energy minimum. It is positive otherwise. It is also checked that the quantity E+P-TS-3nbμ is zero at arbitrary density and temperature. In fact, combining Eqs. (20) and (22) one can easily get P=-Ω0+3nb(μ-μ*), i.e., Ω0=-P+3nb(μ-μ*). Then substituting it into Eq. (25) and combining similar terms, one immediately have

Fig. 4.
The energy (dashed curves) and free energy per baryon (solid curves) of light-quark matter at temperature T=30 MeV. The triangles mark the minimum, while the open circles label the zero pressured points. We notice that the pressure is exactly zero at the free energy minimum, which is a necessary condition for a consistent thermodynamic treatment.
pic
Fig. 5.
The pressure and the free energy per baryon of light-quark matter as functions of the baryon number density at T=50 MeV with parameters D1/2=160 MeV and C=0.6.
pic
E+PTS3nbμ=0. (26)

4 Summary

We have studied the properties of light-quark matter in the equivparticle model with a new quark mass scaling at finite density and temperature. Due to the density and temperature dependence of the quark matter, the chemical potential becomes effective, which ensure the thermodynamic consistency: the pressure at the minimum free energy per baryon is exactly zero.

We extend the quark mass scaling with both quark confinement and first-order perturbative interactions to finite temperature, and apply it to the investigation of light-quark matter. With increasing temperature, the energy per baryon increases while the free energy per baryon decreases. The perturbative interaction increases with density and temperature.

It should be pointed out that the present studies are limited to two-flavor quark matter. Extension to the real three flavor quark matter is a meaningful task, and will be done in the near future.

References
[1] J.F. Xu, G.X. Peng, F. Liu, et al.

Strange matter and strange stars in a thermodynamically self-consistent perturbation model with running coupling and running strange quark mass

. Phys Rev D, 2015, 92: 025025. DOI: 10.1103/PhysRevD.92.025025
Baidu ScholarGoogle Scholar
[2] Y. Nambu, G. Jona-Lasino.

Dynamical model of elementary particles based on an analogy with superconductivity. I

. Phys. Rev, 1961, 122: 345-358. DOI: 10.1103/PhysRev.122.345
Baidu ScholarGoogle Scholar
[3] E.S. Fraga, R.D. Pisarski, J. Schaffner-Bielich.

Small, dense quark stars from perturbative QCD

. Phys Rev D, 2001, 63: 121702(R). DOI: 10.1103/PhysRevD.63.121702
Baidu ScholarGoogle Scholar
[4] G.X. Peng.

Thermodynamic correction to the strong interaction in the perturbative regime

. Europhys Lett, 2005, 72: 69-75. DOI: 10.1209/epl/i2005-10189-8
Baidu ScholarGoogle Scholar
[5] J.F. Xu, G.X. Peng, Z.Y. Lu, et al.

Two-flavor quark matter in the perturbation theory with full thermodynamic consistency

. Sci China-Phys Mech Astron, 2015, 58: 042001. DOI: 10.1007/s11433-014-5599-6
Baidu ScholarGoogle Scholar
[6] S. Plumari, G.F. Bugio, V. Greco V.

Quark matter in neutron stars within the field correlator method

. Phys Rev D, 2013, 88: 083005. DOI: 10.1103/PhysRevD.88.083005
Baidu ScholarGoogle Scholar
[7] R.X. Xu.

Can the age discrepancies of neutron stars be circumvented by an accretion-assisted torque?

Astrophys J, 2003, 596, L75-L78;

Can cold quark matter be solid?

Int J Mod Phys D, 2010, 19: 1437-1446. DOI: 10.1142/S0218271810017767
Baidu ScholarGoogle Scholar
[8] G.S. Khadekar, R. Wanjari.

Geometry of quark and strange quark matter in higher dimensional general relativity

. Int J Theor Phys, 2012, 51: 1408-1415. DOI: 10.1007/s10773-011-1016-3
Baidu ScholarGoogle Scholar
[9] P.K. Sahoo, B. Mishra.

Axially symmetric space-time with strange quark matter attached to string cloud in bimetric theory

. Int J Pure Appl Math, 2013, 82: 87-94. http://ijpam.eu/contents/2013-82-1/8/
Baidu ScholarGoogle Scholar
[10] X.J. Wen.

Color-flavor locked strange quark matter in a strong magnetic field

. Phys Rev D, 2013, 88: 034031. DOI: 10.1103/PhysRevD.88.034031
Baidu ScholarGoogle Scholar
[11] A.A. Isayev, J.J. Yang.

Anisotropic pressure in strange quark matter in the presence of a strong magnetic field

. J Phys G Nucl Part Phys, 2013, 40: 035105. DOI: 10.1088/0954-3899/40/3/035105
Baidu ScholarGoogle Scholar
[12] X.Y. Wang, I.A. Shovkovy.

Bulk viscosity of spin-one color superconducting strange quark matter

. Phys Rev D, 2010, 82: 085007; DOI: 10.1103/PhysRevD.82.085007
M. Huang, I.A. Shovkovy.

Chromomagnetic instability in dense quark matter

. Phys Rev D, 2004, 70: 051501(R). DOI: 10.1103/PhysRevD.70.051501
Baidu ScholarGoogle Scholar
[13] T. Bao, G.Z. Liu, E.G. Zhao, et al.

Self-consistently thermodynamic treatment for strange quark matter in the effective mass bag model

. Eur Phys J A, 2008, 38: 287-293. DOI: 10.1140/epja/i2008-10682-6
Baidu ScholarGoogle Scholar
[14] G.X. Peng, H.C. Chiang, J.J. Yang, et al.

Thermodynamics, strange quark matter, and strange stars

. Phys Rev C, 2000, 61: 015201. DOI: 10.1103/PhysRevC.62.025801
Baidu ScholarGoogle Scholar
[15] G.X. Peng, H.C. Chiang, P.Z. Ning, et al.

Charge and critical density of strange quark matter

. Phys Rev C, 1999, 59: 3452-3454. DOI: 10.1103/PhysRevC.59.3452
Baidu ScholarGoogle Scholar
[16] X.J. Wen, X.H. Zhong, G.X. Peng, et al.

Thermodynamics with density and temperature dependent particle masses and properties of bulk strange quark matter and strangelets

. Phys Rev C, 2005, 72: 015204. DOI: 10.1103/PhysRevC.72.015204
Baidu ScholarGoogle Scholar
[17] J.X. Hou, G.X. Peng, C.J. Xia, et al.

Magnetized strange quark matter in a mass-density-dependent model

. Chin Phys C, 2015, 39: 015101. DOI: 10.1088/1674-1137/39/1/015101
Baidu ScholarGoogle Scholar
[18] S.S. Cui, G.X. Peng, Z.Y. Lu, et al.

Properties of color-flavor locked strange quark matter in an external strong magnetic field

. Nucl Sci Tech, 2015, 26: 040503. DOI: 10.13538/j.1001-8042/nst.26.040503
Baidu ScholarGoogle Scholar
[19] V. Goloviznin, H. Satz.

The refractive properties of the gluon plasma in SU(2) gauge theory

. Z Phys C, 1993, 57: 671-675. DOI: 10.1007/BF01561487
Baidu ScholarGoogle Scholar
[20] A. Peshier, B. Kämpfer, O.P. Pavlenko, et al.

An effective model of the quark-gluon plasma with thermal parton masses

. In: Letessier L, et al, editors. Hot Hadronic Matter. New York: Plenum Press, 1995, 139-142.
Baidu ScholarGoogle Scholar
[21] M.I. Gorenstein, S.N. Yang.

Gluon plasma with a medium-dependent dispersion relation

. Phys Rev D, 1995, 52: 5206-5212. DOI: 10.1103/PhysRevD.52.5206
Baidu ScholarGoogle Scholar
[22] M. Bluhm, B. Kämpfer, G. Soff.

The QCD equation of state near Tc within a quasi-particle model

. Phys Lett B, 2005, 620: 131-136. DOI: 10.1016/j.physletb.2005.05.083
Baidu ScholarGoogle Scholar
[23] V.M. Bannur.

Self-consistent quasiparticle model for quark-gluon plasma

. Phys Rev C, 2007, 75: 044905. DOI: 10.1103/PhysRevC.75.044905
Baidu ScholarGoogle Scholar
[24] F.G. Gardim, F.M. Steffens.

Thermodynamics of quasi-particles at finite chemical potential

. Nucl Phys A, 2009, 825: 222-244. DOI: 10.1016/j.nuclphysa.2009.05.001
Baidu ScholarGoogle Scholar
[25] H. Li, X.L. Luo, H.S. Zong.

Bag model and quark star

. Phys Rev D, 2010, 82: 065017. DOI: 10.1103/PhysRevD.82.065017
L.J. Luo, J.C. Yan, W.M. Sun, et al.

A thermodynamically consistent quasi-particle model without density-dependent infinity of the vacuum zero-point energy

. Eur Phys J C, 2013, 73: 2626. DOI: 10.1140/epjc/s10052-013-2626-0
Baidu ScholarGoogle Scholar
[26] C. Wu, W.L. Qian, R.K. Su.

Improved quark mass density-dependent model with quark and nonlinear scalar field coupling

. Phys Rev C, 2005, 72: 035205. DOI: 10.1103/PhysRevC.72.035205
Baidu ScholarGoogle Scholar
[27] Z.Y. Lu, G.X. Peng, J.F. Xu, et al.

Properties of quark matter in a new quasiparticle model with QCD running coupling

. Sci China-Phys Mech Astron, 2016, 59: 662001. DOI: 10.1007/s11433-015-0524-2
Baidu ScholarGoogle Scholar
[28] G.N. Fowler, S. Raha, R.M. Weiner.

Confinement and phase transitions

. Z Phys C, 1981, 9: 271. DOI: 10.1007/BF01410668
Baidu ScholarGoogle Scholar
[29] S. Chakrabarty, S. Raha, B. Sinha.

Strange quark matter and the mechanism of confinement

. Phys Lett B, 1989, 229: 112-116. DOI: 10.1016/0370-2693(89)90166-4
Baidu ScholarGoogle Scholar
[30] O.G. Benvenuto, G. Lugones.

Strange matter equation of state in the quark mass-density-dependent model

. Phys Rev D, 1995, 51: 1989. DOI: 10.1103/PhysRevD.51.1989;
G. Lugones, O.G. Benvenuto.

Strange matter equation of state and the combustion of nuclear matter into strange matter in the quark mass-density-dependent model at T>0

. Phys Rev D, 1995, 52: 1276. DOI: 10.1103/PhysRevD.52.1276
Baidu ScholarGoogle Scholar
[31] G.X. Peng, H.C. Chiang, B.S. Zhou, et al.

Thermodynamics, strange quark matter, and strange stars

. Phys Rev C, 2000, 62: 025801. DOI: 10.1103/PhysRevC.62.025801
Baidu ScholarGoogle Scholar
[32] G.X. Peng, A. Li, U. Lombardo.

Deconfinement phase transition in hybrid neutron stars from the Brueckner theory with three-body forces and a quark model with chiral mass scaling

. Phys Rev C, 2008, 77: 065807. DOI: 10.1103/PhysRevC.77.065807
Baidu ScholarGoogle Scholar
[33] C.J. Xia, G.X. Peng, S.W. Chen, et al.

Thermodynamic consistency, quark mass scaling, and properties of strange matter

. Phys Rev D, 2014, 89: 105027. DOI: 10.1103/PhysRevD.89.105027
Baidu ScholarGoogle Scholar
[34] G.X. Peng, X.J. Wen, Y.D. Chen.

New solutions for the color-flavor locked strangelets

. Phys Lett B, 2006, 633: 314-318. DOI: 10.1016/j.physletb.2006.01.032
Baidu ScholarGoogle Scholar
[35] X.J. Wen, G.X. Peng, Y.D. Chen.

Charge, strangeness and radius of strangelets

. J Phys G Nucl Part Phys, 2007, 34: 1697. DOI: 10.1088/0954-3899/34/7/010
Baidu ScholarGoogle Scholar
[36] X.P. Zheng, X.W. Liu, M. Kang, et al.

Bulk viscosity of strange quark matter in a density-dependent quark mass model and dissipation of the r mode in strange stars

. Phys Rev C, 2004, 70: 015803. DOI: 10.1103/PhysRevC.70.015803
Baidu ScholarGoogle Scholar
[37] G. Lugones, J.E. Horvath.

Quark-diquark equation of state and compact star structure

. Int J Mod Phys D, 2003, 12: 495. DOI: 10.1142/S0218271803002755
Baidu ScholarGoogle Scholar
[38] P.C. Chu, L.W. Chen.

Quark matter symmetry energy and quark stars

. Astrophys J, 2014, 780: 135. DOI: 10.1088/0004-637X/780/2/135
Baidu ScholarGoogle Scholar
[39] Y. Zhang, R.K. Su.

Quark mass density- and temperature-dependent model for bulk strange quark matter

. Phys Rev C, 2013, 65, 035202. DOI: 10.1103/PhysRevC.65.035202
Baidu ScholarGoogle Scholar
[40] C.J. Xia, S.W. Chen, G.X. Peng.

Properties of strangelets in a new quark mass confinement model with one-gluon-exchange interaction

. Sci China-Phys Mech Astron, 2014, 57: 1304-1310. DOI: 10.1007/s11433-014-5452-y
Baidu ScholarGoogle Scholar
[41] C. Peng, G.X. Peng, C.J. Xia, et al.

Magnetized strange quark matter in the equivparticle model with both confinement and perturbative interactions

. Nucl Sci Tech, 2016, 27: 98. DOI: 10.1007/s41365-016-0095-5
Baidu ScholarGoogle Scholar
[42] G.X. Peng.

Renormalization group dependence of the QCD coupling

. Phys Lett B, 2006, 634: 413. DOI: 10.1016/j.physletb.2006.01.032
Baidu ScholarGoogle Scholar