删除或更新信息,请邮件至freekaoyan#163.com(#换成@)

Strange quark star and the parameter space of the quasi-particle model

本站小编 Free考研考试/2022-01-02

Wen-Hua Cai,, Qing-Wu Wang,College of physics, Sichuan University, Chengdu 610064, China

Received:2021-03-2Revised:2021-06-4Accepted:2021-06-10Online:2021-08-16


Abstract
The properties of strange quark stars are studied within the quasi-particle model. Taking into account chemical equilibrium and charge neutrality, the equation of state (EOS) of $(2+1)$-flavor quark matter is obtained. We illustrate the parameter spaces with constraints from two aspects: one is based on the astronomical results of PSR ${\rm{J}}0740+6620$ and GW 170 817 , and the other is based on the constraints proposed from the theoretical study of a compact star that the EOS must ensure the tidal deformability ${{\rm{\Lambda }}}_{1.4}={190}_{-120}^{+390}$ and support a maximum mass above $1.97\,{M}_{\odot }$. It is found that neither type of constraints can restrict the parameter space of the quasi-particle model in a reliable region and thus we conclude that the low mass compact star cannot be a strange quark star.
Keywords: strange quark;compact star;tidal deformability;quasi-particle model;neutron star


PDF (851KB)MetadataMetricsRelated articlesExportEndNote|Ris|BibtexFavorite
Cite this article
Wen-Hua Cai, Qing-Wu Wang. Strange quark star and the parameter space of the quasi-particle model. Communications in Theoretical Physics, 2021, 73(10): 105202- doi:10.1088/1572-9494/ac09de

1. Introduction

Quantum chromodynamics (QCD), which has been regarded as the theory to govern the strong interaction, predicts that confined hadronic matter will undergo a deconfinement transition to a new phase at large density and extremely low temperature [1, 2]. It is believed that the new phase of nuclear matter, as we call it quark–gluon plasma (QGP), may exist within a new kind of compact star – quark stars [38]. The strange quark matter (SQM) hypothesis indicates that the matter formed by a deconfined u quark, d quark and s quark may be the true ground state of strongly interacting matter [9, 10]. This hypothesis supports the presence of a strange quark star.

In recent years, based on the latest astronomical data, some researchers have argued that many compact stars and pulsars with large mass may be strange quark stars, even including some stars which used to be identified as neutron stars [11]. This naturally raises the question of how to discriminate these strange quark stars from other compact stars. At present, due to the lack of astronomical observation data, it may be achieved through analysis of properties, like the mass–radius relation and tidal deformability, which heavily depends on studying the equation of state (EOS)[12, 13]. Applying an EOS to the Tolman–Oppenheimer–Volkoff (TOV) equation, the relevant properties and features of the star can be obtained. However, at high density and very low temperature, it is difficult to figure out the EOS of cold deconfined quark matter from first principles. Hence, people study the properties of QGP and strange quark stars based on nonperturbative models like the MIT bag model [1417], the Nambu−Jona-Lasinio (NJL) model [1822] and the quasi-particle model [2224].

The quasi-particle model is a kind of phenomenological description of QGP which was obtained by the simulation of lattice QCD [25]. Via the introduction of thermal mass at finite temperature and finite chemical potential, we can treat the system as made up of free quark gas to avoid the complicated calculation of QCD interactions. Thanks to the increasing data from astrophysical observations, it provides a place to test the applicability of an effective model. The latest astrophysical observation of a pulsar, PSR ${\rm{J}}0740+6620$ $(M={2.14}_{-0.09}^{+0.10}\,{M}_{\odot })$ [26], has provided a maximum mass measurement result so far, which is much larger than the result of PSR ${\rm{J}}0348+0432$ $(M=2.01\pm 0.04\,{M}_{\odot })$ [27], and thus the soft EOSs which cannot produce such a massive star are supposed to be eliminated. Moreover, according to the gravitational wave observation GW 170 817 [28], the tidal deformability ${\rm{\Lambda }}$ for 1.4 M star $({{\rm{\Lambda }}}_{1.4})$ needs to be constricted to smaller than 800 in the low-spin prior case. Recently, some significant conclusions have been drawn via the study of quark matter within compact stars, taking into account both the astronomical observations and theoretical simulations [29]. Adopting the speed-of-sound interpolation method, they have succeeded in describing the QCD matter properties of different compact stars with different masses and radii. Their results have been verified by approximately 570 000 EoSs which are built from randomly generated functions. To get such a result, some hypothetical constraints are proposed that the maximum mass should be above 1.97 M and the range of tidal deformability should be revised to ${{\rm{\Lambda }}}_{1.4}={190}_{-120}^{+390}$. Recently, some scholars have even claimed that there may exist a compact star with a mass of 2.6 M [30, 31], which presents great potential to be a candidate for a strange quark star if it is proven to be true. In this paper, from the perspective of strange quark stars, we intend to study the parameter space of the quasi-particle model with some of the latest astrophysical results.

This paper is organized as follows. We first introduce the quasi-particle model at finite chemical potential based on statistical mechanics and thermodynamic equilibrium in section 2. By considering chemical equilibrium and charge neutrality, the EOS of SQM is obtained in section 3. Then, making use of the the EOS, the mass–radius and tidal deformability of a strange quark star are explored. Based on the latest astronomical observations and theoretical research, we study the parameter space of the quasi-particle model. Finally, a short summary of our work is given in section 4.

2. Quasi-particle model at finite chemical potential and zero temperature

The quasi-particle model, which was first proposed to explain the results in lattice gauge theory (LGT) simulations, is widely used to describe the nonperturbative behavior of QCD. Following the works of Bannur [3236], the density of quarks at nonzero temperature is given by$\begin{eqnarray}{\rho }_{i}(T,\mu )=2{N}_{c}\int \displaystyle \frac{{{\rm{d}}}^{3}k}{{\left(2\pi \right)}^{3}}\left(\displaystyle \frac{1}{{{\rm{e}}}^{\tfrac{({\omega }_{i}-\mu )}{T}}+1}+\displaystyle \frac{1}{{{\rm{e}}}^{\tfrac{({\omega }_{i}+\mu )}{T}}+1}\right),\end{eqnarray}$where Nc is the number of colors and subscript i indicates the u, d and s quarks. In addition,$\begin{eqnarray}{\omega }_{i}(T,\mu )=\sqrt{{k}^{2}+{m}_{i}^{2}(T,\mu )},\end{eqnarray}$is the dispersion relation for each kind of quark, where ${m}_{i}(T,\mu )$ stands for the effective mass of different quarks and has the expression$\begin{eqnarray}{m}_{i}^{2}(T,\mu )={\left({m}_{i0}+{m}_{\mathrm{th}}(T,\mu )\right)}^{2}+{m}_{\mathrm{th}}^{2}(T,\mu ).\end{eqnarray}$Here, ${m}_{i0}$ represents the current quark mass and ${m}_{\mathrm{th}}(T,\mu )$ is the thermal mass term that stands for the complicated QCD interaction. In this paper, we take ${m}_{s0}=150\,\mathrm{MeV}$ and ${m}_{u0,d0}\,={m}_{s0}/28.15\approx 5.33$ MeV in the light of [37]. Then, taking the limit $T\to 0$ and computing the integral of equation (1) from zero to Fermi momentum ${k}_{F}=\sqrt{{\mu }^{2}-{m}_{i}{\left(\mu \right)}^{2}}$, we can obtain the number density of quarks$\begin{eqnarray}{\rho }_{i}(\mu )=\displaystyle \frac{{N}_{f}}{3{\pi }^{2}}{\left({\mu }^{2}-{m}_{i}^{2}(\mu )\right)}^{3/2}\theta (\mu -{m}_{i}(\mu )).\end{eqnarray}$The symbol θ in equation (4) stands for the step function. In this case, the thermal mass ${m}_{\mathrm{th}}$ in equation (3) reads$\begin{eqnarray}{m}_{\mathrm{th}}^{2}(\mu )=\displaystyle \frac{{N}_{f}{\mu }^{2}{g}^{2}(\mu )}{18{\pi }^{2}},\end{eqnarray}$in which the symbol Nf is the number of flavors. The effective coupling constant g can be obtained by the two-loop approximation and has the form$\begin{eqnarray}\begin{array}{rcl}{g}^{2}(\mu ) & = & 4\pi {\alpha }_{s}(\mu )=\displaystyle \frac{24{\pi }^{2}}{(33-2{N}_{f})\mathrm{ln}(1.91\mu /2.91\zeta )}\\ & & \times \left[1-\displaystyle \frac{3(153-19{N}_{f})}{{\left(33-2{N}_{f}\right)}^{2}}\displaystyle \frac{\mathrm{ln}(2\mathrm{ln}(1.91\mu /2.91\zeta ))}{\mathrm{ln}(1.91\mu /2.91\zeta )}\right],\end{array}\end{eqnarray}$where ζ is a phenomenological parameter which is related to the nonperturbative effect of QCD [38, 39]. Making use of the physical quantities introduced above, all the physical quantities needed to describe the SQM can be derived through the basic thermodynamic relations.

The relation between number density of quarks ${\rho }_{i}(\mu )$ and chemical potential μ is presented in figure 1. Owing to the step function in the expression of ${\rho }_{i}(\mu )$, it is found that the quark number density vanishes when the chemical potential is below a critical point ${\mu }_{c}$. That is to say, $\mu ={\mu }_{c}$ is a singularity which divides the quark number density into two different regions. This phenomenon is in agreement with the conclusion brought forth in [40], in which researchers pointed out that the existence of some singularity at critical point with zero temperature is a robust and model-independent result based on a universal argument. Similar discussions may also be found in [41, 42].

Figure 1.

New window|Download| PPT slide
Figure 1.Number density ${\rho }_{i}(\mu )$ of u, d and s quarks as functions of chemical potential μ at zero temperature with $\zeta =0.065\,\mathrm{GeV}$, 0.070 GeV and 0.075 GeV respectively.


3. Structure of strange quark stars with the quasi-particle model

Recently, research on strange quark stars has attracted increasing attention. The existing theories suggest that strange quark stars have at least two main channels to be produced. First, a number of them may come into being in the early stage of the universe according to the Big Bang theory. Second, they can also be produced by the phase transition within neutron stars [11]. Therefore, studying strange quark stars does help to understand the phase transition from confined hadron matter to quark matter.

To begin with, there exists chemical equilibrium with the compact stars which is established via weak decay$\begin{eqnarray}d\leftrightarrow u+{e}^{-}+{\bar{\nu }}_{e}\leftrightarrow s.\end{eqnarray}$In consideration of the equilibrium, the electrons and neutrinos have to be included in the system. However, for a stable quark star, we can reasonably assume the neutrinos have enough time to leave the system, which means that there is no need to consider the participation of neutrinos in the chemical equilibrium. To that end, the constraint conditions$\begin{eqnarray}{\mu }_{d}={\mu }_{u}+{\mu }_{e},\end{eqnarray}$and$\begin{eqnarray}{\mu }_{s}={\mu }_{u}+{\mu }_{e},\end{eqnarray}$for the chemical potential of u, d, s quarks and electrons have to be met. Moreover, taking electric charge neutrality into account, the number densities of quarks and electrons should be bound to satisfy the relation$\begin{eqnarray}\displaystyle \frac{2}{3}{\rho }_{u}-(\displaystyle \frac{1}{3}{\rho }_{s}+\displaystyle \frac{1}{3}{\rho }_{d}+{\rho }_{e})=0,\end{eqnarray}$where the electron density ${\rho }_{e}$ is given by ${\rho }_{e}={\mu }_{e}^{3}/(3{\pi }^{2})$ at zero temperature. Then, the relation between baryon number density ${\rho }_{B}=({\rho }_{u}+{\rho }_{d}+{\rho }_{s})/3$ and baryon chemical potential ${\mu }_{B}={\mu }_{u}+{\mu }_{d}+{\mu }_{s}$, as well as the variation of constituents, can be obtained, as shown in figures 2 and 3 respectively. In figure 2, we can see clearly that the baryon number density ${\rho }_{B}({\mu }_{B})$ stays zero when the the baryon chemical potential ${\mu }_{B}$ is smaller than a critical point ${\mu }_{\mathrm{Bc}}$, while in the region of ${\mu }_{B}\gt {\mu }_{\mathrm{Bc}}$, ${\rho }_{B}({\mu }_{B})$ becomes a monotonically increasing function of ${\mu }_{B}$. The proportion of constituents, including u, d and s quarks and electrons, as functions of total baryon number density ${\rho }_{B}$ in weak decays has been shown in figure 3 (the symbol ${\rho }_{\mathrm{sum}}$ stands for the total quark number density and symbol ${\rho }_{0}$ stands for the baryon number density where the s quarks begin to be nonzero). We can see that the graph which represents the proportion of electrons ${\rho }_{e}/{\rho }_{\mathrm{sum}}$ can just be visible after the fraction is multiplied by a factor of 20 . This means that ${\rho }_{e}/{\rho }_{\mathrm{sum}}$ stays in an extremely small range in the SQM system. Meanwhile, the proportion of u quarks is fixed to about 1/3 as a consequence of charge neutrality. Thus, the d quarks occupy the remaining 2/3 before the appearance of s quarks. This outcome agrees with the conclusion in [1], in which the researchers studied the same topic with the NJL model. In addition, we find that the variation of constituents is not associated with ζ. Therefore, we think such behavior of constituents is a model-independent result and only depends on the chemical equilibrium and charge neutrality.

Figure 2.

New window|Download| PPT slide
Figure 2.Baryon number density ${\rho }_{B}$ as a function of baryon chemical potential ${\mu }_{B}$ with $\zeta =0.065\,\mathrm{GeV}$, 0.070 GeV and 0.075 GeV, respectively.


Figure 3.

New window|Download| PPT slide
Figure 3.Proportion of constituent u, d and s quarks and electrons (multiplied by a factor of 20 ) ${\rho }_{i}/{\rho }_{\mathrm{sum}}$ as functions of the baryon number density ${\rho }_{B}/{\rho }_{0}$ (the symbol ${\rho }_{\mathrm{sum}}$ stands for the total quark number density and ${\rho }_{0}$ stands for the baryon number density where the s quarks begin to be nonzero).


The EOS of quark matter at zero temperature can be derived from statistical mechanics and reads [43, 44]$\begin{eqnarray}P(\mu )=P(\mu ){| }_{\mu =0}+{\int }_{0}^{\mu }{\rm{d}}\mu ^{\prime} \rho (\mu ^{\prime} ),\end{eqnarray}$where $P(\mu ){| }_{\mu =0}$ is a negative term. The term $P(\mu ){| }_{\mu =0}$ is the so-called vacuum pressure which is related to the confinement property of QCD. However, due to the lack of comprehensive understanding of QCD interactions, we can neither obtain this term from first principles nor figure out how it is generated. Therefore, we treat the vacuum pressure term as a phenomenological parameter and rewrite it as $P(\mu ){| }_{\mu =0}=-B(B\,\gt 0)$ in this paper. Then, the energy density can be deduced from the thermodynamic relation$\begin{eqnarray}\varepsilon =-P+\sum _{i}{\mu }_{i}{\rho }_{i}.\end{eqnarray}$Combining equation (11) and equation (12) with the discussion taken before, some significant relations can be obtained.

The influence of parameter ζ, parameter B and baryon chemical potential ${\mu }_{B}$ on the EOS of SQM is presented in figures 4 and 5. We can see that the pressure density $P({\mu }_{B})$ and the energy density $\varepsilon ({\mu }_{B})$ have similar behaviors: when ${\mu }_{B}\lt {\mu }_{{Bc}}$, $P({\mu }_{B})$ and $\varepsilon ({\mu }_{B})$ remain constant, and the reason for this phenomenon is the disappearance of the quark number density in this region; once ${\mu }_{B}\gt {\mu }_{{Bc}}$, they become monotonically increasing functions of baryon chemical potential ${\mu }_{B}$. We exhibit the relation between the pressure density and energy density in figure 6. It can be found that the EOS with a larger ζ and a larger B becomes softer. In addition, there is a starting point of the EOS, $(-B,B)$ in the plane of energy–pressure density, which is the consequence of equation (11).

Figure 4.

New window|Download| PPT slide
Figure 4.Pressure density $P({\mu }_{B})$ as a function of baryon number density ${\mu }_{B}$ for different ζ and B.


Figure 5.

New window|Download| PPT slide
Figure 5.Energy density $\varepsilon ({\mu }_{B})$ as a function of baryon number density ${\mu }_{B}$ for different ζ and B.


Figure 6.

New window|Download| PPT slide
Figure 6.Relations between energy density ϵ and pressure density P for different ζ and B.


Utilizing the EOS of SQM, we can get the mass–radius relations (in units of $G=c=1$) by integrating the TOV equations [4547]$\begin{eqnarray}\displaystyle \frac{{\rm{d}}P(r)}{{\rm{d}}r}=-\displaystyle \frac{(\varepsilon +P)(M+4\pi {r}^{3}P)}{r(r-2M)},\end{eqnarray}$and$\begin{eqnarray}\displaystyle \frac{{\rm{d}}M(r)}{{\rm{d}}r}=4\pi {r}^{2}\varepsilon ,\end{eqnarray}$and the result is exhibited in figure 7. As a comparison, we also show the maximum mass constraints required by PSR ${\rm{J}}0740+6620$ $(M={2.14}_{-0.09}^{+0.10}\,{M}_{\odot })$ and PSR ${\rm{J}}0348+0432$ $(M=2.01\pm 0.04\,{M}_{\odot })$, as well as the restriction proposed in [29] in this figure. It is easy to find in figure 7 that a larger maximum mass will be generated with a stiffer EOS. Furthermore, we can find that the EOSs of quasi-particles have the ability to yield a maximum mass larger than the latest observation results. The maximum masses and the corresponding radii of strange quark stars with different ζ and B are calculated, and the results are exhibited in table 1. From this table we can see clearly that the radius of the strange quark star with a mass of near 2 M is close to 12 km. The results are in agreement with the discussions in [4851].

Figure 7.

New window|Download| PPT slide
Figure 7.Mass–radius relations of strange quark stars for different ζ and B with the constraints required by PSR ${\rm{J}}0740+6620$, PSR ${\rm{J}}0348+0432$.



Table 1.
Table 1.Radii R and maximum masses ${M}_{\max }$ of strange quark stars with different ζ and B.
ζ(GeV)${B}^{1/4}$(GeV)R(km)${M}_{\max }({M}_{\odot })$
0.0780.12011.842.00
0.0790.11911.952.01
0.0800.11812.052.02
0.0810.11712.162.03
0.0820.11612.202.04

New window|CSV

In a binary star system, each star will be deformed by the tidal field of its companion because of the presence of tidal gravity. This phenomenon can be described by the physical quantity ‘tidal deformability,' which is defined as the ratio between one star's induced mass quadrupole moment and the tidal field of its companion [47].

In order to calculate the tidal deformability ${\rm{\Lambda }}$, we adopt the relation$\begin{eqnarray}{k}_{2}=\displaystyle \frac{3}{2}{\rm{\Lambda }}{C}^{5},\end{eqnarray}$where $C=M/R$ represents the compactness of a star. Here, the physical quantity k2 is the dimensionless tidal Love number for l=2 which describes the difficulty with which a star can be deformed by an external tidal field, and it can be expressed as$\begin{eqnarray}\begin{array}{rcl}{k}_{2} & = & \displaystyle \frac{8{C}^{2}}{5}{\left(1-2C\right)}^{2}[2+2C(y-1)-y]\\ & & \times \{2C[6-3y+3C(5y-8)]\\ & & +4{C}^{3}[13-11y+C(3y-2)+2{C}^{2}(1+y)]\\ & & +3{\left(1-2C\right)}^{2}[2-y+2C(y-1)]\mathrm{ln}(1-2C){\}}^{-1}.\end{array}\end{eqnarray}$In equation (16), the symbol y is defined as$\begin{eqnarray}y=\displaystyle \frac{R\beta (R)}{H(R)}-\displaystyle \frac{4\pi {R}^{3}{\varepsilon }_{0}}{M},\end{eqnarray}$and is related to metric function $H(r)$ and surface energy density ${\varepsilon }_{0}$. Due to the introduction of negative vacuum pressure, there is a nonzero energy density just around the surface of the star as we can see in figure 6. To derive the metric function, it is necessary to integrate the following differential equations:$\begin{eqnarray}\displaystyle \frac{{\rm{d}}H(r)}{{\rm{d}}r}=\beta ,\end{eqnarray}$$\begin{eqnarray}\begin{array}{rcl}\displaystyle \frac{{\rm{d}}\beta (r)}{{\rm{d}}r} & = & 2{\left(1-2\displaystyle \frac{M}{r}\right)}^{-1}H\left\{-2\pi [5\epsilon +9P+f(\epsilon +P)]\right.\\ & & \left.+\displaystyle \frac{3}{{r}^{2}}+2{\left(1-2\displaystyle \frac{M}{r}\right)}^{-1}{\left(\displaystyle \frac{M}{{r}^{2}}+4\pi {rP}\right)}^{2}\right\}\\ & & +\displaystyle \frac{2\beta }{r}{\left(1-2\displaystyle \frac{M}{r}\right)}^{-1}\left\{-1+\displaystyle \frac{M}{r}+2\pi {r}^{2}(\epsilon -P)\right\},\end{array}\end{eqnarray}$where$\begin{eqnarray}f=\displaystyle \frac{{\rm{d}}\varepsilon }{{\rm{d}}p}.\end{eqnarray}$The integration will start from the center with the expansions $H(r)={a}_{0}{r}^{2}$ and $\beta (r)=2{a}_{0}r$ as the radius $r\to 0$. Since our concern is the ratio of $\beta /H$, we can ignore the value of coefficient a0 and treat it as 1 for simplicity.

Combining the discussions above, the properties of strange quark stars with 1.4 M for different ζ and B can be obtained, as presented in table 2. From this table we can see that the strange quark star is less compact with a smaller ζ and a smaller B for a given mass. This result means that such a star is easier to be deformed by an external tidal field. We illustrate in figure 8 the relations between the tidal deformability for a 1.4 M strange quark star $({{\rm{\Lambda }}}_{1.4})$ and parameters ζ as well as the relations between ${{\rm{\Lambda }}}_{1.4}$ and parameter B. For comparison, we also show the constraint required by GW 170 817 that the ${{\rm{\Lambda }}}_{1.4}$ should be smaller than 800 in the same figure. We can find in figure 8 that not all the parameter settings satisfy the constraint from gravitational wave observations.

Figure 8.

New window|Download| PPT slide
Figure 8.Tidal deformability as function of parameter ζ and bag constant B. (a) Relations between ${{\rm{\Lambda }}}_{1.4}$ and parameter ζ ; (b) relations between ${{\rm{\Lambda }}}_{1.4}$ and parameter B.



Table 2.
Table 2.Properties of strange quark stars with a mass of 1.4 M for different ζ and B, including the compactness $C=M/R$, the Love number k2 as well as the tidal deformability ${{\rm{\Lambda }}}_{1.4}$.
ζ(GeV)${B}^{1/4}$(GeV)$C=M/R$k2${{\rm{\Lambda }}}_{1.4}$
0.1200.0780.1710.187833.682
0.1200.0790.1720.186819.215
0.1200.0800.1730.184787.925
0.1210.0810.1750.181731.478
0.1210.0820.1770.177678.982

New window|CSV

In figure 9, we illustrate the parameter space of the quasi-particle model based on the astronomical observations and theoretical results respectively. In panel (a), we constrain the parameters in view of the PSR ${\rm{J}}0740+6620$ $(M={2.14}_{-0.09}^{+0.10}\,{M}_{\odot })$ and GW 170 817 , and find that there is no coincident region for the feasible area of PSR ${\rm{J}}0740+6620$ and the feasible area of GW 170 817 . Namely, the constraint of GW 170 817 , which requires ${{\rm{\Lambda }}}_{1.4}$ to be smaller than 800 , and the constraint of PSR ${\rm{J}}0740+6620$ $(M={2.14}_{-0.09}^{+0.10}\,{M}_{\odot })$, which requires a maximum mass above 2.14 M, cannot be met simultaneously. Considering the uncertainty of the mass, it can fit to the low boundary of mass as $\zeta \leqslant 0.06\,\mathrm{GeV}$. Similarly, we draw another parameter space in panel (b), which is based on the requirement proposed in [29]. In [29], researchers claimed that the EOS must ensure the tidal deformability ${{\rm{\Lambda }}}_{1.4}={190}_{-120}^{+390}$ and support a maximum mass above 1.97 M. It is clear to see from (b) that the two feasible areas do not coincide when ζ is larger than about 0.03 GeV, and in the region of $\zeta \lt 0.03\,\mathrm{GeV}$, there is a small coincident area for the two constraints, which almost vanishes. However, according to [52], the quasi-particle model we use will work better in the large ζ region in order to fit the LGT data; as for the area of small ζ, the results may be unreliable to some extent, let alone the coincident region which is very small compared to the whole parameter space.

Figure 9.

New window|Download| PPT slide
Figure 9.Parameter space of the quasi-particle model based on observation results of PSR ${\rm{J}}0740+6620$ $(M={2.14}_{-0.09}^{+0.10}\,{M}_{\odot })$ and GW 170 817 and based on constraints proposed in [29], respectively.


In figure 9 the parameter space of the quasi-particle model cannot meet the the constraints of PSR ${\rm{J}}0740+6620$ $(M={2.14}_{-0.09}^{+0.10}\,{M}_{\odot })$ and GW 170 817 simultaneously. Meanwhile, the parameter space also cannot restrict the requirement proposed in [29] in a reliable region. Up to now, increasing evidence has indicated that the compact stars of large masses may be strange quark stars [5356]. On the basis of the results in figure 9, we think it is probably improper to regard a compact star with a mass of 1.4 M as strange quark star. Naturally, we assume that the properties of compact stars with small mass (including 1.4 M) look more like neutron stars formed by hadronic matter rather than strange quark stars. This hypothesis is consistent with the conclusion of [29], in which the researchers claimed that they found quark matter exists within compact stars with 2 M but is not present within compact stars with $1.4\,{M}_{\odot }$.

4. Summary

In this paper, we have studied the properties of strange quark stars with the quasi-particle model. Taking the chemical equilibrium of the weak interaction and charge neutrality into account, we have obtained the baryon number density and the variation of constituents for different ζ. We have found that the proportions of quarks are irrelevant with ζ. With the introduction of parameter B for the vacuum pressure, we have established the EOS of SQM. With proper choices of B and ζ, it is possible to generate a quark star with a maximum mass larger than $2.14\,{M}_{\odot }$, which matches the results of PSR ${\rm{J}}0740+6620$ $(M={2.14}_{-0.09}^{+0.10}\,{M}_{\odot })$ and PSR ${\rm{J}}0348+0432$ $(M=2.01\pm 0.04\,{M}_{\odot })$. We have calculated the properties of a strange quark star with mass of $1.4\,{M}_{\odot }$, including the compactness C, the Love number k2 and the tidal deformability ${{\rm{\Lambda }}}_{1.4}$. It is found that a softer EOS corresponds to a more compact quark star, which is more difficult to be deformed.

Finally, we have illustrated the parameter space of the quasi-particle model based on astronomical observations and theoretical suggestions. It is found that the parameter space cannot meet the constraint of GW 170 817 which requires ${{\rm{\Lambda }}}_{1.4}$ smaller than 800 , and the constraint of PSR ${\rm{J}}0740+6620$ $(M={2.14}_{-0.09}^{+0.10}\,{M}_{\odot })$ which requires a maximum mass of at least $2.14\,{M}_{\odot }$, simultaneously. Similarly, the model also cannot satisfy the constraints proposed in [29], in which the researchers claimed that the EOS must ensure the tidal deformability ${{\rm{\Lambda }}}_{1.4}={190}_{-120}^{+390}$ and support a maximum mass above $1.97\,{M}_{\odot }$.

Early studies on neutron star models are in good agreement with the observed data of low mass stars. However, the EOSs are relatively soft, so it is difficult to get a maximum mass larger than 2 ${M}_{\odot }$. We can obtain large mass objects based on the EOS of SQM, but cannot obtain correct Λ data for stars with small mass simultaneously. Therefore, in order to meet the Λ data from the present astronomical observations, it is not feasible for the present three flavored quark model. Of course, our conclusion should be understood as model dependent. There may be other possibilities, such as thinking of a pulsar as a three flavored gigantic nucleus which is called a strangeon star [57]. In figure 1 of [57], with ${\rm{\Lambda }}(1.4)=400$, the maximum mass of strangeon stars can be larger than $3{M}_{\odot }$. Therefore, our calculations as compared with the quark-cluster model of a strangeon star may indicate that the coupling between quarks in a strange star is still very strong so that quarks are localized/confined in strangeons, as is in the case of nucleons for a nucleus. A recent work assuming the SQM is in the color–flavor locked phase has found that the color superconductivity gap is poorly constrained by those observed global properties of a $1.4{M}_{\odot }$ star [58]. As such, it is suggested to use a hadronic EOS in exploring the properties of low mass compact stars while using a pure quark EOS or hybrid EOS in studying compact stars with larger mass.

Acknowledgments

This work is supported by the Cultivating Plan of Characteristic Direction of Science (2020SCUNL209).


Reference By original order
By published year
By cited within times
By Impact factor

Buballa M 2005 Phys. Rep. 407 205
DOI:10.1016/j.physrep.2004.11.004 [Cited within: 2]

Luo X F Xu N 2017 Nucl. Sci. Tech. 28 112
DOI:10.1007/s41365-017-0257-0 [Cited within: 1]

Ivanenko D Kurdgelaidze D F 1969 Lett. Nuovo Cimento 2 13
DOI:10.1007/BF02753988 [Cited within: 1]

Itoh N 1970 Prog. Theor. Phys. 44 291
DOI:10.1143/PTP.44.291

Iwamoto N 1980 Phys. Rev. Lett. 44 1637
DOI:10.1103/PhysRevLett.44.1637

Bodmer A R 1971 Phys. Rev. D 4 1601
DOI:10.1103/PhysRevD.4.1601

Haensel P Zdunik J L Schaeffer R 1986 Astron. Astrophys. 160 121


Alcock C Farhi E Olinto A 1986 Astrophys. J. 310 261
DOI:10.1086/164679 [Cited within: 1]

Terazawa H 1989 J. Phys. Soc. Jpn. 58 3555
DOI:10.1143/JPSJ.58.3555 [Cited within: 1]

Witten E 1984 Phys. Rev. D 30 272
DOI:10.1103/PhysRevD.30.272 [Cited within: 1]

Kuerban A Geng J J Huang Y F Zong H S Gong H 2019 arXiv:1908.11191
[Cited within: 2]

Ozel F 2006 Nature 441 1115
DOI:10.1038/nature04858 [Cited within: 1]

Alford M et al. 2007 Nature 445 E7
DOI:10.1038/nature05582 [Cited within: 1]

Chodos A Jaffe R L Johnson K Thorn C B Weisskopf V F 1974 Phys. Rev. D 9 3471
DOI:10.1103/PhysRevD.9.3471 [Cited within: 1]

Alford M Braby M Paris M Reddy S 2005 Astrophys. J. 629 969
DOI:10.1086/430902

Alcock C Farhi E Olinto A 1986 Astrophys. J. 310 261
DOI:10.1086/164679

Zhou E-P Zhou X Li A 2018 Phys. Rev. D 97 083015
DOI:10.1103/PhysRevD.97.083015 [Cited within: 1]

Wang Q-W Shi C Zong H-S 2019 Phys. Rev. D 100 123003
DOI:10.1103/PhysRevD.100.123003 [Cited within: 1]

Wang Q-W Xiq Y Zong H 2018 Mod. Phys. Lett. A 33 1850232
DOI:10.1142/S0217732318502322

Fan Z-Y Fan W-K Wang Q-W Zong H-S 2017 Mod. Phys. Lett. A 32 1750107
DOI:10.1142/S0217732317501073

Menezes D P Providncia C Melrose D B 2006 J. Phys. G 32 1081
DOI:10.1088/0954-3899/32/8/001

Peshier A Kampfer B Soff G 2000 Phys. Rev. C 61 045203
DOI:10.1103/PhysRevC.61.045203 [Cited within: 2]

Szabo K K Toth A I 2003 J. High Energy Phys. 06 008
DOI:10.1088/1126-6708/2003/06/008

Plumari S Alberico W M Greco V Ratti C 2011 Phys. Rev. D 84 094004
DOI:10.1103/PhysRevD.84.094004 [Cited within: 1]

Ma H-H et al. 2018 arXiv:1804.06797
[Cited within: 1]

Cromartie H T et al. 2019 Nat. Astron. 4 72
DOI:10.1038/s41550-019-0880-2 [Cited within: 1]

Antoniadis J et al. 2013 Science 340 1233232
DOI:10.1126/science.1233232 [Cited within: 1]

Abbott B P et al. 2017 Phys. Rev. Lett. 119 161101
DOI:10.1103/PhysRevLett.119.161101 [Cited within: 1]

Annala E Gorda T Kurkela A Nattila J Vuorinen A 2020 Nat. Phys. 16 907
DOI:10.1038/s41567-020-0914-9 [Cited within: 8]

Fattoyev F J Horowitz C J Piekarewicz J Reed B 2020 Phys. Rev. C 102 065805 arXiv:2007.03799
[Cited within: 1]

Tan H Noronha-Hostler J Yunes N 2020 Phys. Rev. Lett. 125 261104 arXiv:2006.16296
[Cited within: 1]

Bannur V M 2007 Eur. Phys. J. C 50 629
DOI:10.1140/epjc/s10052-007-0233-7 [Cited within: 1]

Bannur V M 2007 Phys. Lett. B 647 271
DOI:10.1016/j.physletb.2007.02.030

Bannur V M 2007 Phys. Rev. C 75 044905
DOI:10.1103/PhysRevC.75.044905

Bannur V M 2007 JHEP 09 046
DOI:10.1088/1126-6708/2007/09/046

Bannur V M 2008 Phys. Rev. C 78 045206
DOI:10.1103/PhysRevC.78.045206 [Cited within: 1]

Ma H-H Qian W-L 2018 Braz. J. Phys. 48 160
DOI:10.1007/s13538-017-0551-7 [Cited within: 1]

Rebhan A Romatschke P 2003 Phys. Rev. D 68 025022
DOI:10.1103/PhysRevD.68.025022 [Cited within: 1]

Schneider R A 2003 arXiv:hep-ph/0303104
[Cited within: 1]

Halasz M A Jackson A D Shrock R E Stephanov M A Verbaarschot J J M 1998 Phys. Rev. D 58 096007
DOI:10.1103/PhysRevD.58.096007 [Cited within: 1]

Tian Y-L Yan Y Li H Luo X-L Zong H-S 2012 Phys. Rev. D 85 045009
DOI:10.1103/PhysRevD.85.045009 [Cited within: 1]

Simji P 2020 Int. J. Mod. Phys. A 35 2050064
DOI:10.1142/S0217751X20500645 [Cited within: 1]

He M Feng H T Sun W M Zong H S 2007 J. Phys. G 34 2655
DOI:10.1088/0954-3899/34/12/010 [Cited within: 1]

Zong H-S Sun W-M 2008 Phys. Rev. D 78 054001
DOI:10.1103/PhysRevD.78.054001 [Cited within: 1]

Hinderer T 2008 Astrophys. J. 677 1216
DOI:10.1086/533487 [Cited within: 1]

Damour T Nagar A 2009 Phys. Rev. D 80 084035
DOI:10.1103/PhysRevD.80.084035

Yagi K Yunes N 2013 Science 341 365
DOI:10.1126/science.1236462 [Cited within: 2]

Li B-L Cui Z-F Yu Z-H Yan Y An S Zong H-S 2019 Phys. Rev. D 99 043001
DOI:10.1103/PhysRevD.99.043001 [Cited within: 1]

Zong H-S Sun W-M 2008 Phys. Rev. D 78 054001
DOI:10.1103/PhysRevD.78.054001

Annala E Gorda T Kurkela A Vuorinen A 2018 Phys. Rev. Lett. 120 172703
DOI:10.1103/PhysRevLett.120.172703

Wang Q-W Shi C Yan Y Zong H-S 2019 arXiv:1912.02312
[Cited within: 1]

Bannur V M 2012 Int. J. Mod. Phys. E 21 1250090
DOI:10.1142/S0218301312500905 [Cited within: 1]

Li C-M Zuo S-Y Yan Y Zhao Y-P Wang F Huang Y-F Zong H-S 2020 Phys. Rev. D 101 063023
DOI:10.1103/PhysRevD.101.063023 [Cited within: 1]

Jokela N Jarvinen M Remes J 2019 JHEP 03 041
DOI:10.1007/JHEP03(2019)041

Pinkanjanarod S Burikham P 2020 arXiv:2007.10615


Otto K Oertel M Schaefer B-J 2020 arXiv:2007.07394 [hep-ph]
[Cited within: 1]

Lai X Zhou E Xu R 2019 Eur. Phys. J. A 55 60
DOI:10.1140/epja/i2019-12720-8 [Cited within: 2]

Li A Jiang J-L Tang S-P Miao Z-Q Zhou E-P Xu R-X 2020 Eur. Phys. J. Spec. Top. 229 3629 arXiv:2007.07394
[Cited within: 1]

相关话题/Strange quark parameter