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

\catchline

NEUTRON STARS AS DENSE LIQUID DROP AT EQUILIBRIUM WITHIN THE EFFECTIVE SURFACE APPROXIMATION

A.G. Magner Nuclear Theory Department, Institute for Nuclear Research, 03028 Kyiv, Ukraine
Cyclotron Institute, Texas A&M University, College Station, Texas 77843, USA
[email protected]
   S.P. Maydanyuk Nuclear Processes Department, Institute for Nuclear Research, 03028 Kyiv, Ukraine
Institute of Modern Physics, Chinese Academy of Sciences, Lanzhou, 730000, China
[email protected]
   A. Bonasera Cyclotron Institute, Texas A&M University, College Station, Texas 77843, USA
[email protected]
   H. Zheng School of Physics and Information Technology, Shaanxi Normal University,Xi’an 710119, China
[email protected]
   A.I. Levon Nuclear Reactions Department, Institute for Nuclear Research, 03028 Kyiv, Ukraine
[email protected]
   T.M. Depastas Cyclotron Institute, Texas A&M University, College Station, Texas 77843, USA
[email protected]
   U.V. Grygoriev Nuclear Theory Department, Institute for Nuclear Research, 03028 Kyiv, Ukraine
[email protected]
(Day Month Year; Day Month Year)
Abstract

The macroscopic model is formulated for a neutron star (NS) as a perfect liquid drop at the equilibrium. We use the leptodermic approximation a/R1a/R\ll 1, where aa is the crust thickness of the effective NS surface (ES), and RR is the mean radius of the ES curvature. Within the approximate Schwarzschild metric solution to the general relativity theory equations for the spherically symmetric systems, the macroscopic gravitation is taken into account in terms of the total separation particle energy and incompressibility. Density distribution ρ\rho across the ES in the normal direction to the ES was obtained analytically for a general form of the energy density (ρ)\mathcal{E}(\rho). For the typical crust thickness, and effective radius, one finds the leading expression for the density ρ\rho. NS masses are analytically calculated as a sum of the volume and surface terms, taking into account the radial curvature of the metric space, in reasonable agreement with the recently measured masses for several neutron stars. We derive the simple macroscopic equation of state (EoS) with the surface correction. The analytical and numerical solutions to Tolman-Oppenheimer-Volkoff equations for the pressure are in good agreement with the volume part of our EoS.

keywords:
Neutron stars; dense liquid drop; Schwarzschild metric; energy density; effective surface; equation of state.
{history}
\ccode

PACS number: 21.65.Mn,26.60.Gj

1 INTRODUCTION

R.C. Tolman suggested [1] first to study the simplest model for a neutron star (NS) considering it as a dense liquid-matter drop at its equilibrium under the gravitational, nuclear and other realistic forces; see also his book [2], chapt. 7, sect. 96. Within this model, the Einstein-Gilbert equations of the General Relativistic Theory (GRT) for the spherical symmetry case has been reduced to the three independent equations for four unknown quantities, namely, two parameters, λ\lambda and ν\nu, of the Schwarzschild metric (see also Ref. 3, the chapters 11 and 12)

ds2=eλdr2r2dθ2r2sin2θdϕ2+eνc2dt2,{\rm d}s^{2}=-e^{\lambda}{\rm d}r^{2}-r^{2}{\rm d}\theta^{2}-r^{2}\hbox{sin}^{2}\theta{\rm d}\phi^{2}+e^{\nu}c^{2}{\rm d}t^{2}, (1)

and the pressure PP, and energy density \mathcal{E}. To get the complete system of equations Tolman suggested also to derive independently the equation of state (EoS), =(ρ)\mathcal{E}=\mathcal{E}(\rho), or P=P(ρ)P=P(\rho). The EoS can be found from the condition of a static equilibrium for a liquid-matter drop under gravitational, and nuclear (and Coulomb) forces, which all should be taken into account by the simplified GRT equations for the gravitational field. Following Tolman’s ideas, Oppenheimer and Volkoff[4] derived the simple (TOV) equations by using essentially the macroscopic properties of the system [Refs. 3 (chapt. 12) and 5 (chapt. 1)] of the system as a perfect liquid drop at equilibrium. As shown in Ref. 2, one can derive the analytical solution for the pressure PP as function of the radial coordinate. So far, the TOV equations [4] are considered with the equation of state, =(ρ)\mathcal{E}=\mathcal{E}(\rho), which was obtained independently of the macroscopic assumptions used in the TOV derivations. For instance, the EoS with a polytropic expression for the pressure[6, 7] P=P(ρ)P=P(\rho) as function of the density ρ\rho is assumed to be similar to that for a particle gas system with fitted parameters. The gradient terms are usually neglected in the energy density, and then, no equilibrium for a finite gas system with no external fields. However, the pressure PP was calculated as function of the radial rr coordinate by using the TOV equations for another dense liquid-drop system. Such a system, including the gravitation, can be described mainly at a stable equilibrium by short-range forces, in contrast to a gas matter with long-range inter-particle interaction. Then, the NS mass MM was obtained numerically as function of the NS radius RR, M=M(R)M=M(R), with important restrictions due to those of the NS radius. In any case, we should emphasize the importance of the gradient terms in the energy density (ρ)\mathcal{E}(\rho) for a macroscopic condition of the system equilibrium, in spite of a relative small NS crust thickness. As a hint to these conclusions, see many applications of the TOV equations, e.g., in Refs. 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18. Concerning the relation of the nuclear and neutron liquid-drop models (LDMs) to NS properties, we should also mention the work[19] by Baym, Bethe, and Pethick. They discuss the liquid matter drop with the leptodermic property having a sharp decrease of the particle number density in a relatively small edge region considered as its surface. Clear specific definitions and complete updated results for the energy density with density gradient (surface) terms and for equations of the infinite matter state with many inter-particle forces in the non-relativistic and relativistic cases can be found in the recent review Ref. 18.

Refer to caption   Refer to caption


Figure 1: Left: A qualitative plot for the deformed ES with the local coordinate system ξ,η\xi,\eta, where ξ\xi is the axis perpendicular, and η\eta is that parallel to the ES. We show also a full diffuse surface thickness 2a2a. Right: Particle number density, ρ\rho, in units of the saturation value, ρ¯\overline{\rho}, as function of the variable x=ξ/a=(rR)/ax=\xi/a=(r-R)/a for the NS in a simple compressed LDM at the stable equilibrium. Solid line is related to the asymmetric solution, Eq. (32), for β=γ=0\beta=\gamma=0 (line “1”). Dashed line presents the same but for the Wilets symmetric solution, Eq. (33) (line “2”). Parameters, for example: the effective radius R=10R=10 km, diffuseness of the NS crust a=1a=1 km. The dots ES1 and ES2 show the ES at the effective radius RR for the solid, Eq. (32), and dashed, Eq. (33), lines, respectively.

Taking Tolman’s ideas, one can try first to extend the EoS to those for a dense macroscopic[3, 5] system of particles. In this leptodermic system, the particle density ρ\rho is function of the radial coordinate with exponentially decreasing behavior from an almost constant saturation value inside of the dense system to that through the effective surface (ES) of NS in a relatively small crust range aa; see Fig. 1. The ES is defined as the points of spatial coordinates with a maximum density gradient. To obtain the analytical solutions for the particle density and EoS, we use the leptodermic approximation[3, 2, 4, 18, 21, 22, 23, 24, 25, 26, 27, 19, 20, 17, 6], a/R1a/R\ll 1, where RR is the (curvature) radius of the ES.

Within this effective surface approximation (ESA), a/R1a/R\ll 1, simple and accurate solutions of many nuclear and liquid-drop problems involving the particle number density distributions were obtained for nuclei[28, 29, 30, 31, 32, 33, 34, 35]. The ESA exploits the property of saturation of the nuclear particle density ρ\rho inside of the system, which is a characteristic feature of dense systems as molecular systems111 For the one-dimensional and more complicate dense molecular (e.g., liquid-drop) systems, van der Waals (vdW) suggested the phenomenological capillary theory[36] which predicted the results for the particle number density ρ\rho and surface tension coefficients σ\sigma. These results are similar to those obtained later in Refs. 28, 29; see also comments below., liquid drops, nuclei, and presumably, NSs. The realistic energy-density distribution is minimal at a certain saturation density of particles (nucleons, neutrons, or nuclei) corresponding approximately to the infinite matter[37]. As a result, relatively narrow edge region exists in finite nuclei or NS (crust) in which the density drops sharply from its almost central value to zero. We assume here that the part inside of the system far from the ES can be changed a little (saturation property of the dense system as hydrostatic liquid drop, nucleus or NS in the final evolution state).

The coordinate system related to the effective surface is defined in such a way that one of the spatial coordinates (ξ\xi) is the distance from the given point to the ES; see Fig. 1, and A. This non-linear coordinate system is conveniently used in the region of nuclear and NS edges. They allow for an easy extraction of relatively large terms in the density distribution equations for the variation equilibrium condition. This condition means that the variation of the total energy EE over the density ρ\rho is zero under the constraints which fix some integrals of motion beyond the energy EE by the Lagrange method. The Lagrange multipliers are determined by these constraints within the local energy-density theory, in particular, the extended Thomas-Fermi (ETF) approach from nuclear physics[39, 38]. Neglecting the other smaller-order perpendicular- and all parallel-to-ES contributions, sum of such terms leads to a simple one-dimensional equation (in special local coordinates with the coordinate normal-to-surface ξ\xi); see Fig. 1.

Such an equation mainly determines approximately the density distribution across the diffused surface layer of the relatively small order of the ratio of the diffuseness parameter aa to the (mean curvature) ES radius RR, a/R1a/R\ll 1, for sufficiently heavy systems. Notice that within this manuscript, as in Refs. 28, 29, 30, 31, 32, 33, 34, 35, the “diffuseness parameter”, “ the crust range”, and “the thickness of the system edge” have the same meaning for neutron stars. A small parameter, a/Ra/R, of the expansion within the ESA can be used for analytical solving the variational problem for the minimum of the system energy with constraints for a fixed particle number, and other integrals of motion, such as angular momentum, quadrupole deformation, etc. When this edge distribution of the density is known, the leading static and dynamic density distributions which correspond to diffused surface conditions can be easily constructed. To do so, one has to determine the dynamics of the effective surface which is coupled to the volume dynamics of the density by certain LDM boundary conditions[40, 31, 41]. This ESA approach is based on the catastrophe theory for solving differential equations with a small parameter of the order of a/Ra/R as the coefficient in front of the high order derivatives in normal to the ES direction[42]. A relatively large change of the density ρ\rho on a small distance aa with respect to the curvature radius RR takes place for the liquid-matter drop (nuclei, water drops, neutron stars). Inside of such dense systems, the density ρ\rho is changed slightly around a constant saturation density ρ¯\overline{\rho}. Therefore, one obtains essential effects of the surface capillary pressure. Another important idea of Tolman is that we should consider as simpler as possible the neutron star in terms of the liquid drop at a static equilibrium under the stability condition, i.e., having a minimum of the total energy under constraints mentioned above.

The accuracy of the ESA was checked in Ref. 32 for the case of nuclear physics by comparing the results with the existing nuclear theories like Hartree-Fock[43] (HF) and extended Thomas-Fermi[39, 38] approaches, based on the Skyrme forces [43, 44, 45, 46, 47, 39, 50, 48, 9, 49], but for the simplest case without spin-orbit and asymmetry terms of the energy density functional. The direct variational principle for finding numerically the parameters of the tested particle density functions in simple forms of the Woods-Saxon-like in Ref. 38 or their powers (Ref. 51) were applied by using the realistic Skyrme energy functional[9, 49]. The main focus in Ref. 51 aimed to the surface corrections to the nuclear symmetry energy of spherical nuclei, see also recently published variational ETF approach[52, 53]. The extension of the ES approach to the nuclear isotopic symmetry and spin-orbit interaction has been done in Refs. 33, 34, 35. The Swiatecki derivative terms of the symmetry energy for heavy nuclei[54, 55, 56, 57, 58, 59, 60, 61] were taken into account within the ESA in Ref. 35. The discussions of the progress in nuclear physics and astrophysics within the relativistic local density approach, can be found in reviews Refs. 18, 62; see also Refs. 13, 63, 64. Notice that the simplest nuclear Thomas-Fermi approach (without gradient terms in the energy density) was applied for neutron stars in Ref. 65.

In the present work, we extend the ES approximation of Refs. 32, 33, 34, 35, to the spherical neutron stars by the method working also for deformed systems. In Sect. 2, we present the basic local energy-density formalism taking into account particle density gradients which are responsible for surface terms in finite macroscopic dense systems. The particle number density solutions inside of the nucleus are obtained analytically within the ESA expansion in a small leptodermic parameter a/Ra/R at leading order in Sect. 3. The leading density distributions for the density ρ\rho at the lowest order in a relatively small diffuse surface-layer size, a/R1a/R\ll 1, are obtained analytically in Sect. 4. The NS mass with the surface correction is derived in Sect. 5 taking into account the Schwarzschild metric. The surface energy in terms of the tension coefficients of the vdW macroscopic capillary theory[42, 40, 31, 41] is obtained analytically through vdW-Skyrme forces parameters in Sect. 6. Equation of state in the form P=P(ρ)P=P(\rho), for the system surface corrections, is derived through the ES at the same leading order in Sect. 7. The TOV approach is presented in Sect. 8 for a step-like particle number density. The results of our calculations and comparison of the macroscopic and some semi-microscopic calculations of the EoS are discussed in Sect. 9. The main results are summarized in Sect. 10. Some details of the mathematical textbook relations will be shown in A.

2 LOCAL ENERGY DENSITY AND CONSTRAINTS

The total energy EE for static problems can be written as

E=d𝒱[ρ(𝐫)],E=\int\hbox{d}\mathcal{V}\;\mathcal{E}[\rho({\bf r})], (2)

where (ρ){\cal E}(\rho) is the energy density[32, 35],

(ρ)=𝒜(ρ)+(ρ)(ρ)2.\mathcal{E}\left(\rho\right)=\mathcal{A}(\rho)+\mathcal{B}(\rho)\left(\nabla\rho\right)^{2}. (3)

The integration is carried out over the volume of a system, d𝒱=eλ/2d𝐫\hbox{d}\mathcal{V}=e^{\lambda/2}\hbox{d}{\bf r}, where λ\lambda is the coefficient of the Schwarzchild metric (1) for a spherical system. As shown in Ref. 3, the multiplier eλ/2e^{\lambda/2} takes into account the gravitational defect of the NS mass. In Eq. (3), 𝒜(ρ)\mathcal{A}(\rho) and (ρ)\mathcal{B}(\rho) are smooth functions of the density ρ\rho which are coefficients in expansion of the energy density over gradients of ρ\rho. A non-gradient part 𝒜(ρ)\mathcal{A}(\rho) of the energy density \mathcal{E} can be written as

𝒜=bVρ+ε(ρ)+mρΦ(ρ),\mathcal{A}=-b_{V}\rho+\varepsilon(\rho)+m\rho\Phi(\rho), (4)

where bVb_{V} is the non-gravitational energy component for the separation of particle from the matter, mm is the particle mass, and Φ\Phi is the macroscopic gravitational potential determined in more details below. The second and third terms take into account the non-gravitational (like nuclear) and gravitational contributions into the incompressibility. In order to define properly other quantities in Eq. (4), ε(ρ)\varepsilon(\rho) and Φ(ρ)\Phi(\rho), we now use the condition for a minimum of the energy per particle, 𝒲\mathcal{W}, at a stable equilibrium,

(d𝒲dρ)ρ=ρ¯=0,𝒲=ρ.\left(\frac{\hbox{d}\mathcal{W}}{\hbox{d}\rho}\right)_{\rho=\overline{\rho}}=0~{},\quad\mathcal{W}=\frac{\mathcal{E}}{\rho}~{}. (5)

Near the saturation value, ρρ¯\rho\rightarrow\overline{\rho}, one has 𝒲=𝒜/ρ\mathcal{W}=\mathcal{A}/\rho. Therefore, there is no linear terms in expansion of 𝒜\mathcal{A}, Eq. (4), over powers of the difference ρρ¯\rho-\overline{\rho} near the saturation value for an isolated system including the gravitation. For instance, for ε(ρ)\varepsilon(\rho) in Eq. (4), one can write, generally speaking, for the considered dense system (similarly as in Refs. 32, 33),

ε(ρ)=K18ρ¯2ρ(ρρ¯)2,\varepsilon(\rho)=\frac{K}{18\overline{\rho}^{2}}~{}\rho\left(\rho-\overline{\rho}\right)^{2}, (6)

where KK is the non-gravitational part of the incompressibility modulus, e.g., due to the Skyrme nuclear interaction (for nuclear matter K200K\sim 200 MeV). In Eq. (4), as mentioned above, Φ(ρ)\Phi(\rho) is the main statistically (macroscopically) averaged part of the gravitational potential. In what follows, we may restrict ourselves by only quadratic terms in expansion of the smooth function 𝒜(ρ)\mathcal{A}(\rho) due to the saturation property of the NS inside of the dense finite drop of the matter.

As well known[2, 3], the Einstein-Gilbert GRT equations in the non-relativistic limit and/or for a week gravitational field can be transformed to the continuity Poisson equation for the gravitational potential Φ\Phi. As shown in Ref. 3, in this limit, g00=eνg_{00}=e^{\nu} of a more general Schwarzschild metric, Eq. (1), can be presented as eν1+ν=1+2Φ/c2e^{\nu}\approx 1+\nu=1+2\Phi/c^{2}, where Φ\Phi is the solution of the Poisson equation[3, 7] ν=2Φ/c2\nu=2\Phi/c^{2}. We will use a more general definition for the gravitational potential, Φ=(c2/2)lng00=(c2/2)ν\Phi=(c^{2}/2)\ln g_{00}=(c^{2}/2)\nu. Instead of the Newtonian limit expansion, for the statistically averaged potential Φ\Phi, we will use another macroscopic approximation within the leptodermic approach. Therefore, after such a macroscopic averaging, Φ\Phi can be considered as function of the radial coordinate rr through the particle number density ρ=ρ(r)\rho=\rho(r) in the form of expansion of Φ(ρ)\Phi(\rho) over powers of ρρ¯\rho-\overline{\rho} near the saturation density ρ¯\overline{\rho} for leptodermic systems. This is similar to the macroscopic Coulomb potential in nuclear physics, considered up to the residual inter-particle interaction[29, 30, 33] but accounting now for second order terms in ρρ¯\rho-\overline{\rho}. Up to second order, one can take into account the macroscopically averaged gravitational contribution to the incompressibility modulus KK,

Φ=Φ0+Φ1(ρρ¯)+(1/2)Φ2(ρρ¯)2,\Phi=\Phi_{0}+\Phi_{1}(\rho-\overline{\rho})+(1/2)\Phi_{2}(\rho-\overline{\rho})^{2}, (7)

where Φ0=Φ(ρ¯)\Phi_{0}=\Phi(\overline{\rho}), Φn=nΦ(ρ¯)/ρn\Phi_{n}=\partial^{n}\Phi(\overline{\rho})/\partial\rho^{n} (n=1,2) are the derivatives of the gravitational potential Φ(ρ)\Phi(\rho) at the saturation density ρ¯\overline{\rho}. According to Eq. (4) for the non-gradient part 𝒜\mathcal{A} of the energy density \mathcal{E}, and the saturation condition (5), one finds for the zero constant Φ1\Phi_{1} in the expansion (7) of the gravitational potential Φ(ρ)\Phi(\rho). With this condition and the expansion (7), one obtains

Φ=Φ0+(1/2)Φ2(ρρ¯)2.\Phi=\Phi_{0}+(1/2)\Phi_{2}(\rho-\overline{\rho})^{2}~{}. (8)

Therefore, for 𝒜(ρ)\mathcal{A}(\rho), Eq. (4), we arrive at

𝒜=bV(G)ρ+εG(ρ),bV(G)=bVmΦ0.\mathcal{A}=-b^{(G)}_{V}\rho+\varepsilon_{G}(\rho)~{},\quad b^{(G)}_{V}=b_{V}-m\Phi_{0}~{}. (9)

It was convenient to introduce the two new quantities, bV(G)b^{(G)}_{V} is the total separation energy per particle, and εG(ρ)\varepsilon_{G}(\rho) is the total second-order energy density component, both modified by the gravitational field,

εG(ρ)=ε(ρ)+m2Φ2(ρρ¯)2=KG18ρ¯2ρ(ρρ¯)2.\varepsilon_{G}(\rho)=\varepsilon(\rho)+\frac{m}{2}\Phi_{2}(\rho-\overline{\rho})^{2}=\frac{K_{G}}{18\overline{\rho}^{2}}~{}\rho\left(\rho-\overline{\rho}\right)^{2}~{}. (10)

Here, KGK_{G} is the total incompressibility modulus modified by the gravitational field (KG>0K_{G}>0),

KG=K+9mρ¯2Φ2.K_{G}=K+9m\overline{\rho}^{2}\Phi_{2}~{}. (11)

Notice that our second order approximation for the gravitational potential Φ\Phi expansion, Eq. (8), agrees with the energy density presentation up to second order gradients, Eq. (3). The latter is, in turn, rather general for all known Skyrme forces in nuclear physics taking into account the volume and surface terms. Thus, we account for the contribution of a strong gravitational field in terms of separation energy bV(G)b^{(G)}_{V}, Eq. (9), and the total incompressibility KGK_{G}, Eq. (11), which will be agreed with the Schwarzschild metric space curvature within the ES approximation. As we will see, this second-order potential term is consistently related with the density gradient squared term of the energy density (ρ)\mathcal{E}(\rho), Eq. (3), by the equilibrium equation at leading order of the leptodermic parameter a/Ra/R.

The coefficient (ρ)\mathcal{B}(\rho) in front of the gradient squared term in Eq. (3) is given by

(ρ)=𝒞+𝒟ρ+Γρ.\mathcal{B}(\rho)=\mathcal{C}+\mathcal{D}\rho+\frac{\Gamma}{\rho}~{}. (12)

These terms are associated with the nuclear Skyrme interaction. First term is related to the interaction term which is a main reason of the diffuse surface thickness for a non-rotating NS. In this sense it has a more general meaning including the main effective interaction in a dense molecular system, studied by van der Waals[36] (vdW); see also the footnote on page 4. Therefore, we will call this component as the vdW-Skyrme interaction. The second term is coming from the spin-orbit interaction. This interaction might be important, for instance, for any dense rotating liquid-drop systems with a sharp surface edge, i.e., with large density-gradient terms in the surface region for a finite leptodermic systems; see also the same general arguments for using the ETF approach in Refs. 38, 39. The last term is a gradient correction to the kinetic energy (2\hbar^{2} correction of the nuclear kinetic energy in the ETF approach[39, 38]). It is introduced here for a comparison of the dense and gas systems. Thus, the forms (9) for 𝒜\mathcal{A} and (12) for \mathcal{B} are rather general among the simplest analytical solutions for dense finite systems.

We should add the constraint for the variational procedure to get equations for the static equilibrium. For non-rotated one-component NS system we have to fix the particle number NN,

N=d𝒱ρ(𝐫).N=\int\hbox{d}\mathcal{V}\rho({\bf r})~{}. (13)

The integration is carried out over the volume occupied by the gravitational mass and determined by the GRT metric. Therefore, introducing the chemical potential μ\mu as the Lagrange multiplier, one obtains from Eqs. (2) and (3) the equation for the equilibrium,

δδρ𝒜ρρ(ρ)22Δρ=μ,\frac{\delta\mathcal{E}}{\delta\rho}\equiv\frac{\partial\mathcal{A}}{\partial\rho}-\frac{\partial\mathcal{B}}{\partial\rho}\left(\nabla\rho\right)^{2}-2\mathcal{B}\Delta\rho=\mu~{}, (14)

where 𝒜(ρ)\mathcal{A}(\rho) and (ρ)\mathcal{B}(\rho) are given by Eqs. (9) and (12), respectively. Equation (13) determines the chemical potential μ\mu in terms of the particle number NN. For nuclear liquid drop, one has two equations related to the two constraints for the fixed neutron and proton numbers (Refs. 33, 34, 35). They determine two (neutron and proton) chemical potentials. The Coulomb interaction can be taken into account for a proton part of the nucleus through the Coulomb potential; see Refs. 30, 33. The terms like Δρ\propto\Delta\rho with constant of the proportionality are omitted because they do not affect the equilibrium density because of their disappearance from the variational equations for the total energy owing to the well-known Ostragradsky-Gauss theorem. High order derivative terms in the energy per particle /ρ\mathcal{E}/\rho [see Eq. (3) for \mathcal{E}], are neglected for simplicity in analogy of the GRT, Ref. 2. Equation (3) for \mathcal{E} overlaps most of the Skyrme forces[9, 49]. As function of a local particle density ρ\rho, the (ρ)\mathcal{E}(\rho) corresponds to a saturation condition, Eq. (5). The most remarkable in this form is that the energy density (ρ)\mathcal{E}(\rho), Eq. (3), is a sum of the three terms related to the incompressible constant, bV(G)-b^{(G)}_{V}, Eq. (9), and the compressible energy, (ρρ¯)2\propto(\rho-\overline{\rho})^{2}, both including the gravitational components, Eq. (8), and the surface gradient terms. In nuclear physics, one has, instead of the gravitational term but similarly, the Coulomb potential (even without quadratic terms). Similarly, the symmetry energy, and surface (2\nabla^{2}) terms can be taken into account in nuclear physics and astrophysics. Within the nuclear ETF approach, the terms being proportional to Γ\Gamma of the gradient part in (3) comes from the 2\hbar^{2} correction to the TF kinetic energy density[38], Γ=2/18m\Gamma=\hbar^{2}/18m, where mm is the nucleon mass. In the gradient squared part, Eq. (12), the 𝒞\mathcal{C} constant term is typical for the potential component of the Skyrme energy density functional and 𝒟\mathcal{D} is the constant of the spin-orbit term[39] in the ETF model in nuclear physics. For a nucleus, the spin-orbit coefficient 𝒟\mathcal{D} is given by 𝒟=(9m/162)W02\mathcal{D}=-(9m/16\hbar^{2})W_{0}^{2}, where W0=100130W_{0}=100-130 MeV fm5 is the nuclear spin-orbit constant; see Refs. 39, 9, 49. The Coulomb part of the nuclear energy density (3) is considered similarly as suggested in Refs. 29, 30. Meaning of all terms in the energy density (3) will be specified more below.

For the spherical and deformed system of NN particles, we may find the equilibrium particle density ρ\rho from the variational problem for the energy functional (2) with respect to the variations of the density δρ\delta\rho, which obey the constraints in the form:

N=d𝒱ρ(𝐫),Q=d𝒱ρ(𝐫)q(𝐫).N=\int\hbox{d}\mathcal{V}\;\rho({\bf r}),\quad Q=\int\hbox{d}\mathcal{V}\;\rho({\bf r})\;q({\bf r})~{}. (15)

The last constraint fixes a certain deformation parameter. The function q(𝐫)q({\bf r}), for instance, can be a multipole moment, or it can be chosen in such a way that Q determines the distance between the centers of masses of the two halves of the stretched nucleus for fission problems[66]. Thus, for the energy functional (2), (3) and the constraints (15), one finds the same Lagrange variational equation (14) but μ\mu would be equal to the sum of the chemical potential and qμQq\mu_{Q}, μμ+qμQ\mu\Rightarrow\mu+q\mu_{Q}. Lagrange multipliers μ\mu, and μQ\mu_{Q} are determined by the two constraints (15). Formally, we may consider the same equation (14) for spherical and deformed systems taking into account qq in μ\mu.

3 VARIATIONAL EQUATION IN THE SYSTEM VOLUME

In the system volume, the terms of Eq. (14) containing derivatives of ρ\rho are small. These derivatives in a normal direction become large near the nuclear edge. For a rather wide class of deformed shapes of the dense finite system, as well as for the axially-symmetric, in particular, spherical one, we may assume that the thickness aa of a system edge is small as compared with its mean curvature radius RR, considering a/Ra/R as a small parameter. In this respect, we define the effective surface as the points of maximum of the gradient of a particle density ρ\nabla\rho, as shown in Fig. 1 by dots in the right panel. Locally near a point of the ES, one may introduce the local coordinate system ξ,η\xi,\eta, where ξ\xi is normal to the ES, and η\eta presents two other orthogonal coordinates; see for instance, as shown in Fig. 1 and A. In particular, for the spherical coordinates, η\eta can be two spherical angle coordinates. See A, also for simple geometric relations for the non-linear ξ,η\xi,\eta coordinates. It naturally appears the mean curvature HH in terms of a small leptodermic parameter of the ES approximation, aHaH; see Eq. (79). For heavy enough nuclei near the spherical shape with the effective radius R=r0A1/3R=r_{0}A^{1/3}, one has aH=a/RA1/31aH=a/R\sim A^{-1/3}\ll 1 because a/r01a/r_{0}\sim 1 for realistic nuclear parameters, r0=(4πρ¯/3)1/31.14r_{0}=(4\pi\overline{\rho}/3)^{-1/3}\approx 1.14 fm at the density of infinite nuclear matter, ρ¯=0.16\overline{\rho}=0.16 fm-3, and the typical diffuseness parameter a0.8a\approx 0.8 fm, see below more precise definition for the diffuseness parameter through the decrement of decreasing of the exponential asymptotes of particle density. For NSs, it is well known that the typical surface diffuseness (the NS crust) a1a\approx 1 km, and the mean effective radius R10R\approx 10 km; see, e.g., Ref. 20. Therefore, the ratio a/Ra/R can be also considered as a small parameter, and the leptodermic approximation, a/R1a/R\ll 1, can be used too.

The largest terms in Eq. (14) within the region of a sharp density descent are the second-order derivative of particle density ρ\rho in the ξ\xi direction, normal to the ES, d2ρ/dξ2\hbox{d}^{2}\rho/\hbox{d}\xi^{2}, and its derivative square, (dρ/dξ)2(\hbox{d}\rho/\hbox{d}\xi)^{2}, both of the order of (ρ¯/a)2(R/a)2(\overline{\rho}/a)^{2}\propto(R/a)^{2}; see the expression for the Laplacian and gradient in ξ,η\xi,\eta coordinates in A. Our approach is based on expansion in the same small parameter aHaH (for a nucleus, aH=a/R1/kFRA1/3aH=a/R\sim 1/k_{F}R\sim A^{-1/3}, where kFk_{F} is the Fermi momentum in units of \hbar). This leptodermic approach is used in the liquid-drop and the extended Thomas-Fermi approach. Following Refs. 29, 31, 32, 33, 34, 35, we shall call this statistical222The ETF in nuclear physics is the statistical semiclassical approach because for convergence of the expansion in \hbar of the partition function we have first to average statistically it removing strong (e.g., shell) oscillations, and thus, get a smooth behavior of the partition function[38]. semiclassical ETF approach as the ES approximation. We have to evaluate also the derivatives of the particle density and gravitational terms in the energy density (ρ)\mathcal{E}(\rho), Eq. (3), and Lagrange equation (14). For simplicity, the gravitational terms [see Eqs.(8) and (14)] are assumed to be of the same order as other non-gradient terms. Notice that according to these estimations, the transformation of radial derivatives due to the Schwarzschild radial curvature can be taken into account through the leptodermic parameter a/Ra/R because they are important near the ES. The gravitational corrections are in contrast to those due to the Coulomb potential for the nuclear liquid drop, where the Coulomb corrections to a constant are assumed to be negligible as a/Ra/R. The latter corrections[30] are of the relative order (e2/c)(c/bVr0)(Z2/A4/3)/2 < 0.2\sim(e^{2}/\hbar c)(\hbar c/b_{V}r_{0})(Z^{2}/A^{4/3})/2\hbox{\kern 1.00006pt\lower 2.58334pt\hbox{$\sim$} \kern-11.19997pt\raise 2.58334pt\hbox{$<$} \kern 1.00006pt}0.2.

As function (ρ)\mathcal{E}(\rho) obeys the saturation property (5), in the system volume at ξ < ξvol\xi\hbox{\kern 1.00006pt\lower 2.58334pt\hbox{$\sim$} \kern-11.19997pt\raise 2.58334pt\hbox{$<$} \kern 1.00006pt}\xi_{vol} (ξ<0\xi<0 and |ξ|a|\xi|\gg a), where ρρ¯\rho-\overline{\rho} is small asymptotically far from the ES, one can expand the non-gradient function 𝒜\mathcal{A} of the energy density (ρ)\mathcal{E}(\rho), Eq. (3), up to second order including the compression term of the second order in powers of ρρ¯\rho-\overline{\rho}, 𝒜=𝒜V\mathcal{A}=\mathcal{A}_{V}; see Eqs. (9), (10), and (6),

(ρ)𝒜V=bV(G)ρ+KG18ρ¯2ρ(ρρ¯)2.\mathcal{E}(\rho)\rightarrow\mathcal{A}_{V}=-b^{(G)}_{V}\rho+\frac{K_{G}}{18\overline{\rho}^{2}}~{}\rho(\rho-\overline{\rho})^{2}~{}. (16)

The zero and second order terms of expansion of the gravitational potential, Eq. (8), are included in the zero and second order terms through constants bV(G)b^{(G)}_{V} [Eq. (9)] and KGK_{G} [Eq. (11)], respectively.

Introducing the dimensionless quantities for convenience to exclude the transformation of units,

y=ρρ¯,x=ξa,y=\frac{\rho}{\overline{\rho}},\quad x=\frac{\xi}{a}~{}, (17)

one can present Eq. (16) in the following way:

εG(ρ)𝒜V(ρ)+bV(G)ρ=KGρ¯18ϵ(y),\varepsilon_{G}(\rho)\equiv\mathcal{A}_{V}(\rho)+b^{(G)}_{V}\rho=\frac{K_{G}\overline{\rho}}{18}~{}\epsilon(y), (18)

where

ϵ(y)=y(1y)2.\epsilon(y)=y(1-y)^{2}~{}. (19)

Discarded terms are of the order of ((ρρ¯)/ρ¯)3(a/R)3((\rho-\overline{\rho})/\overline{\rho})^{3}\sim(a/R)^{3}. Neglecting gradient terms in Eq. (14) in the system volume and using the approximation (8) for the gravitational potential Φ\Phi, for simplicity, up to quadratic terms over ρρ¯\rho-\overline{\rho}, one can reduce equation (14) as

εGρKG9(ρρ¯)=,\frac{\partial\varepsilon_{G}}{\partial\rho}\equiv\frac{K_{G}}{9}~{}(\rho-\overline{\rho})=\mathcal{M}~{}, (20)

where \mathcal{M} is an integration constant. Solving this equation with respect to the particle number density ρ\rho, one obtains

ρ=ρ¯(1+9KG),=μ+bV(G).\rho=\overline{\rho}\left(1+\frac{9\mathcal{M}}{K_{G}}\right),\quad\mathcal{M}=\mu+b^{(G)}_{V}~{}. (21)

Therefore, \mathcal{M} is the surface (capillary) correction to the leading component of the chemical potential μ\mu [introduced above as the Lagrange multiplier, determined through the constraint (13)]. As seen from Eq. (21), one finds the relatively small surface correction, 9/KGa/R9\mathcal{M}/K_{G}\sim a/R, to the saturation constant value ρ¯\overline{\rho}; see Refs.  29, 30, 31, 33, 34, 35. Therefore, \mathcal{M} is also the capillary pressure correction; see also Ref. 42.

4 DENSITY NEAR THE EFFECTIVE SURFACE

In this section, we will analytically solve equation (14) for the distribution of the particle density ρ\rho through the ES at leading order in small parameter a/Ra/R. Equation (14) is a typical differential equation in the catastrophe theory. In such differential equations, a small coefficient (of the order of a/Ra/R in our case) appears in front of the highest order derivative. These terms become important because of the product of this small coefficient (of the order of a/Ra/R) and large (of the order of R/aR/a and high) can be of zero order in this leptodermic parameter. At the leading order, we need to keep only leading terms in small parameter a/Ra/R in the Lagrange equation (14). These are the second order derivatives ρ′′\rho^{\prime\prime}, and first-order derivative squares (ρ)2(\rho^{\prime})^{2}, over their ξ\xi variable, according to Eqs. (80) and (77), ρ=ρ/ξ\rho^{\prime}=\partial\rho/\partial\xi. We may develop some iteration procedure to find the solutions with improved precision in a/Ra/R in terms of the solutions of the leading order. As mentioned above, the transformation of radial derivatives due to a smooth Schwarzschild radial curvature for the spherical case can be taken into account near the ES through the re-definition of the leptodermic parameter a/Ra/R.

Multiplying Eq. (14) by the derivative ρ/ξ\partial\rho/\partial\xi, from Eqs. (14), (9), and (12) at leading order, one obtains

ddξ[(𝒞+𝒟ρ+Γ4ρ)(dρdξ)2εG(ρ)]=0,\frac{\hbox{d}}{\hbox{d}\xi}\left[\left(\mathcal{C}+\mathcal{D}\rho+\frac{\Gamma}{4\rho}\right)\;\left(\frac{\hbox{d}\rho}{\hbox{d}\xi}\right)^{2}-\varepsilon_{G}\left(\rho\right)\right]=0~{}, (22)

where

εG(ρ)=ρ¯KG18ϵ(y),\varepsilon_{G}(\rho)=\frac{\overline{\rho}K_{G}}{18}\epsilon(y), (23)

and ϵ(y)\epsilon(y) is a given function of yy, which is determined by a short- and long-range inter-particle interaction. In particular, one can use a simple quadratic approximation (19) far from the critical point where KG=0K_{G}=0. In Eq. (22), the term with the coefficient 𝒞\mathcal{C} in parentheses in front of the derivative squared of the density ρ\rho is the main term, e.g., as related to the vdW capillary theory. In particular, it is the main term, associated with the Skyrme interaction, with respect to the spin-orbit term 𝒟ρ\mathcal{D}\rho and, moreover, the gradient correction Γ/(4ρ)\Gamma/(4\rho) to the kinetic energy of the system in nuclear physics. Let us first assume the same for the NS liquid drop. Therefore, aa should be determined through the constant 𝒞\mathcal{C}. Otherwise, we have to include average of whole coefficient in front of the derivative square term in Eq. (22).

With employing the boundary conditions ρ0\rho\to 0 and ρ0\rho^{\prime}\to 0 for ξ\xi\to\infty, one can easily integrate equation (22). Finally, we come to a simple ordinary 1st order differential equation for ρ\rho, depending only on the normal-to-ES coordinate ξ\xi, at leading order in a small parameter a/Ra/R, that is

dρdξ=εG(ρ)𝒞+𝒟ρ+Γ/4ρ.\frac{\hbox{d}\rho}{\hbox{d}\xi}=-\sqrt{\frac{\varepsilon_{G}(\rho)}{\mathcal{C}+\mathcal{D}\rho+\Gamma/4\rho}}~{}. (24)

Introducing several dimensionless quantities,

y=ρρ¯,x=ξa,ϵ=18εGKGρ¯,y=\frac{\rho}{\overline{\rho}},\quad x=\frac{\xi}{a},\quad\epsilon=\frac{18\varepsilon_{G}}{K_{G}\overline{\rho}}~{}, (25)

one can relate the crust thickness aa to the vdW interaction constant 𝒞\mathcal{C} by

a=18𝒞ρ¯KG.a=\sqrt{18\mathcal{C}\frac{\overline{\rho}}{K_{G}}}~{}. (26)

Using these definitions, one can present Eq. (24) in the following dimensionless form:

dydx=yϵ(y)y+βy2+γ,\frac{\hbox{d}y}{\hbox{d}x}=-\sqrt{\frac{y\epsilon(y)}{y+\beta y^{2}+\gamma}}~{}, (27)

where

β=𝒟ρ¯𝒞,γ=Γ4ρ¯𝒞.\beta=\frac{\mathcal{D}\;\overline{\rho}}{\mathcal{C}},\quad\gamma=\frac{\Gamma}{4\overline{\rho}\;\mathcal{C}}~{}. (28)

From the asymptotes of the explicit analytical expressions for the particle density y(x)y(x) at large xx, one may see a typical behavior, y(x)exy(x)\propto e^{-x}, where x=ξ/ax=\xi/a, that is a reason to call aa, Eq. (26), as the crust diffuseness parameter. Thus, we may check that aa has exactly the same meaning as that introduced above in terms of the small parameter a/Ra/R. As noted above, in the derivation of Eq. (27) within leading order of the ES expansion, we reduced the second order differential equation (14) to the first order one by integrating with using the standard boundary conditions y0y\to 0 and y/ξ0\partial y/\partial\xi\to 0 and yϵ(y)0y\epsilon(y)\to 0 for the definition of the ϵ(y)\epsilon(y) asymptotes. Finally, all parameters of the energy density in Eq. (3) are reduced within this leading order ES approximation to the only two dimensionless constants; see Eq. (28). For nuclear physics,

β=27W02128πkFr06εF𝒜,γ=πεFr0527𝒜(kFr0)2,\beta=-\frac{27W_{0}^{2}}{128\pi\;k_{F}\;r_{0}^{6}\;\varepsilon_{F}\;{\cal A}},\quad\gamma=\frac{\pi\varepsilon_{F}\;r_{0}^{5}}{27{\cal A}\left(k_{F}\;r_{0}\right)^{2}}, (29)

where εF=2kF2/2m37\varepsilon_{F}=\hbar^{2}k^{2}_{F}/2m\approx 37 MeV is the Fermi energy for the Fermi momentum, kF=(3π2ρ¯/2)1/3k_{F}=(3\pi^{2}\;\overline{\rho}/2)^{1/3} (in units of \hbar). We are coming to the realistic nuclear data for the infinite-matter particle density, ρ¯=0.16\overline{\rho}=0.16 fm-3.

Single boundary condition which we need for the unique solution of the first order differential equation can be easily found within the leading ES approximation from the definition of the ES as the points of maximal values of the derivative, ρ/ξ\partial\rho/\partial\xi at ξ=0\xi=0, namely, y′′(0)=0y^{\prime\prime}(0)=0 at x=0x=0. Differentiating equation (27) over xx, one obtains the algebraic equation for the position of the ES, y=y0y=y_{0} at x=0x=0,

(γβy2)ϵ(y0)+y0(y0+βy02+γ)dϵ(y0)dy=0,(\gamma-\beta y^{2})\epsilon(y_{0})+y_{0}(y_{0}+\beta y_{0}^{2}+\gamma)\frac{\hbox{d}\epsilon(y_{0})}{\hbox{d}y}=0~{}, (30)

defined by function ϵ(y)\epsilon(y); see, e.g., Eq. (19).

For any function ϵ(y)\epsilon(y), Eq. (27) can be easily integrated. The integration constant is determined by the boundary condition (30). The leading-order solution of Eq. (27) can be found explicitly in the inverse form:

x=y0ydww+βw2+γwϵ(w).x=-\int_{y_{0}}^{y}\hbox{d}w\sqrt{\frac{w+\beta w^{2}+\gamma}{w\epsilon(w)}}~{}. (31)

For the expression (19) for ϵ(y)\epsilon(y), this integral can be calculated analytically in terms of the elementary functions for any parameters β\beta and γ\gamma within the condition a/R1a/R\ll 1. Note that the solution (31) to equation (27) is considered at leading order in small parameter a/Ra/R. This solution satisfies asymptotically the condition of its matching with the volume result, Eq. (21), of the same order at the point xx corresponding ξξvol\xi\approx\xi_{vol} defined in Section 3, y1y\to 1 exponentially for xx\to-\infty.

Let us consider the examples of simple solutions x=x(y)x=x(y), Eq. (31), of Eq. (27) with the boundary condition, Eq. (30), for the quadratic expression ϵ(y)=y(1y)2\epsilon(y)=y(1-y)^{2}.

(i) For γ=0\gamma=0 and β=0\beta=0, neglecting the gradient correction to the kinetic energy density (Γ=0\Gamma=0 in Eq. (12)), and spin-orbit (𝒟=0\mathcal{D}=0 in Eq.(12)) terms, one finally obtains[32, 33]

y(x)=tanh2[(xx0)/2],x<x0=2arctanh(1/3).y(x)=\tanh^{2}[(x-x_{0})/2],\quad x<x_{0}=2{\rm arctanh}(1/\sqrt{3})~{}. (32)

This solution is related to the gradient squared term due to the main nuclear Skyrme, or molecular van der Waals interaction term [𝒞0\mathcal{C}\neq 0 in Eq.(12)]. It is very asymmetric with respect to the ES value, y0=y(0)=1/3y_{0}=y(0)=1/3; see Fig. 1(b).

(ii) Keeping only Γ0\Gamma\neq 0 term in Eq. (24), but neglected gradient terms of the vdW-Skyrme (𝒞=0\mathcal{C}=0) and spin-orbit (𝒟=0\mathcal{D}=0) interaction, one obtains333Dimensionless equation in this case is dy/dx=(yϵ(y))1/2\hbox{d}y/\hbox{d}x=-(y\epsilon(y))^{1/2} for a=Γ[9/(8KG)]1/2a=\Gamma[9/(8K_{G})]^{1/2}. Equation for the ES, x=0x=0, ϵ(y0)+y0dϵ(y0)/dy=0\epsilon(y_{0})+y_{0}\hbox{d}\epsilon(y_{0})/\hbox{d}y=0, has the solution y0=1/2y_{0}=1/2. Integrating analytically the equation for y(x)y(x) in this case, one obtains the expression (33). the Wilets solution[28], derived early for the semi-infinite system,

y(x)=1/(1+ex).y(x)=1/(1+e^{x})~{}. (33)

This solution is symmetric with respect to the ES because of y(0)=1/2y(0)=1/2. This Fermi-gas form of the solutions is used for the variational problems within the ETF model, in contrast to the dense liquid-drop solution (32), related to the vdW-Skyrme interaction; see Fig. 1(b).

Figure 1(b) shows the particle density, ρ(r)=ρ¯y((rR)/a)\rho(r)=\overline{\rho}y((r-R)/a), as function of the radial coordinate rr for typical parameters, the thickness of the NS diffused layer (crust) a=1a=1 km, and the effective radius R=10R=10 km, in units of the saturation value ρ¯\overline{\rho}. We compare the two dimensionless solutions for the main asymmetric particle density ρ(r)/ρ¯\rho(r)/\overline{\rho}. One of them (solid line) is related to the corresponding vdW-Skyrme interaction “1”; see Eq. (32) for the universal dimensionless particle density of a dense liquid-drop, y=ρ/ρ¯y=\rho/\overline{\rho}, as function of x=(rR)/ax=(r-R)/a. Another limit case corresponds to the symmetric Wilets solution based on Eq. (33) for zero constants 𝒞\mathcal{C} and 𝒟\mathcal{D} in Eq. (12) for \mathcal{B} (only the gradient correction to the kinetic energy is taken into account, i.e. for a gas system). This solution “2” is often used for the variational version of this problem. These two curves are very different in the surface layer, especially outside of the system far away from the NS, and almost the same inside of the NS. Notice that so far the effective surface method can be applied for the deformed NS under the condition of smallness of crust thickness aa over the effective NS radius RR determined by the mean surface curvature HH, aH=a/R1aH=a/R\ll 1; see A. However, in the following sections, our macroscopic method will be applied for the simplest solved analytically case of the spherical symmetry.

5 NS mass surface correction

Let us derive the NS mass M within our macroscopic ESA approach,

M=mρd𝒱,M=m\int\rho\hbox{d}\mathcal{V}~{}, (34)

where mm is the particle (nucleon, nucleus, or molecule) mass. The integration is carried out over the spatial volume,

d𝒱=J(r)r2drsinθdθdφ,J(r)=eλ/2.\hbox{d}\mathcal{V}=J(r)r^{2}\hbox{d}r~{}\hbox{sin}\theta~{}\hbox{d}\theta\hbox{d}\varphi~{},\quad J(r)=e^{\lambda/2}~{}. (35)

This element of the spatial volume is associated with the Schwarzschild interior spherical coordinate system (see Refs. 2, 3),

ds2=dr21r2/RSM2r2dθ2r2sin2θdϕ2+[ASMBSM1r2/RSM2]2dt2,{\rm d}s^{2}=-\frac{{\rm d}r^{2}}{1-r^{2}/R^{2}_{\rm SM}}-r^{2}{\rm d}\theta^{2}-r^{2}\hbox{sin}^{2}\theta{\rm d}\phi^{2}+\left[A_{\rm SM}-B_{\rm SM}\sqrt{1-r^{2}/R^{2}_{\rm SM}}\right]^{2}{\rm d}t^{2}~{}, (36)

where RSMR_{\rm SM}, ASMA_{\rm SM}, and BSMB_{\rm SM} are some constants; see Ref. 2, and more details below. Here and below we will use the subscript SM to relate the quantities to the Schwarzschild metric, Eq. (36). The Schwarzschild radius RSMR_{\rm SM} is given by[2, 3]

RSM=(8πG03c4)1/2,R_{\rm SM}=\left(\frac{8\pi G\mathcal{E}_{0}}{3c^{4}}\right)^{-1/2}, (37)

where GG is the gravitational constant, 0=𝒜(ρ¯)=bV(G)ρ¯\mathcal{E}_{0}=\mathcal{A}(\overline{\rho})=-b^{(G)}_{V}\overline{\rho}, Eq. (9). Then, according to Eqs. (35) and (36), for the radial part of the Jacobian factor JJ in this transformation, one has

J(r)=(1r2RSM2)1/2.J(r)=\left(1-\frac{r^{2}}{R^{2}_{\rm SM}}\right)^{-1/2}. (38)

The Jacobian JJ takes approximately into account the gravitational defect mass[3].

Adding and subtracting ρ¯\overline{\rho} in the integrand of Eq. (34), one can identically rewrite the NS mass MM in the following way:

M=MV+MS.M=M_{V}+M_{S}~{}. (39)

Here, the first volume component of the NS mass, MVM_{V}, is given by

MV=4πmρ¯0Rr2dr(1r2/RSM)1/2=4πmρ¯RSM3f(RRSM),M_{V}=4\pi m\overline{\rho}\int_{0}^{R}\frac{r^{2}{\rm d}r}{\left(1-r^{2}/R_{\rm SM}\right)^{1/2}}=4\pi m\overline{\rho}R_{\rm SM}^{3}f\left(\frac{R}{R_{\rm SM}}\right)~{}, (40)

where

f(z)=12[arcsin(z)z1z2]f(z)=\frac{1}{2}\left[\arcsin(z)-z\sqrt{1-z^{2}}\right] (41)

for 0<z<10<z<1. The second term, MSM_{S}, in Eq. (39) is the surface component of the NS mass,

MS=mSR2(ρρ¯)J(r)r2dr,M_{S}=\frac{mS}{R^{2}}\int(\rho-\overline{\rho})J(r)r^{2}\hbox{d}r~{}, (42)

where SS is the surface area of the spherical system, S=4πR2S=4\pi R^{2}. The integral over the radial coordinate rr is taken effectively over a small diffuse crust region of the order of aa, where ρρ¯\rho-\overline{\rho} essentially differs from zero. A smooth Jacobian J(r)J(r), Eq. (38), multiplied by r2r^{2}, for rr far away from the Schwarzschild radius RSMR_{\rm SM}, can be taken approximately off the integral in Eq. (42) over the radial coordinate rr at the ES r=Rr=R, J(r)J(R)J(r)\approx J(R). Using the differential equation (24) for the density ρ\rho at leading order in a/Ra/R in the local coordinate system ξ,η\xi,\eta (A), one can transform the integration variable ξ\xi to ρ\rho, dr=dξ=(dξ/dρ)dρ\hbox{d}r=\hbox{d}\xi=(\hbox{d}\xi/\hbox{d}\rho)\hbox{d}\rho. Thus, the surface correction, MSM_{S}, Eq. (42), in terms of the dimensionless density y=ρ/ρ¯y=\rho/\overline{\rho}, is resulted in

MSmaSρ¯J(R)01(1y)y+βy2+γyϵ(y)dy,M_{S}\approx-maS\overline{\rho}J(R)\int_{0}^{1}\frac{(1-y)\sqrt{y+\beta y^{2}+\gamma}}{\sqrt{y\epsilon(y)}}\hbox{d}y, (43)

where J(R)J(R) is given by Eq. (38) for J(r)J(r) at r=Rr=R, and ϵ(y)\epsilon(y) can be parameterized by using the vdW-Skyrme interaction model but with NS parameters. For the simplified (quadratic) expression of ϵ(y)\epsilon(y), Eq. (19), the integral in Eq. (43) can be taken analytically in terms of the elementary functions. In particular, for the vdW-Skyrme case (i), β=γ=0\beta=\gamma=0, Eqs. (32) and (38) at r=Rr=R, one explicitly finds from Eq. (43)

MS8πmρ¯R2a(1R2/RSM2)1/2.M_{S}\approx-8\pi m\overline{\rho}R^{2}a\left(1-R^{2}/R^{2}_{\rm SM}\right)^{-1/2}. (44)

Finally, we arrive from Eqs. (39), (40), and (44) at the simple result in the case (i):

M=MV[12aR2RSM3f(R/RSM)(1R2/RSM2)1/2],M=M_{V}\left[1-\frac{2aR^{2}}{R^{3}_{\rm SM}f\left(R/R_{\rm SM}\right)\left(1-R^{2}/R^{2}_{\rm SM}\right)^{1/2}}\right], (45)

where MVM_{V}, Eq. (40), is the volume part of the total NS mass MM [see Eq. (39)], i.e., the mass MM at a=0a=0, and f(z)f(z) is given by Eq. (41).

For completeness, one also can present the NS mass with the surface correction in the Wilets case (ii), 𝒞=𝒟=0\mathcal{C}=\mathcal{D}=0, Eq. (33),

M=MV[16aγR2RSM3f(R/RSM)(1R2/RSM2)1/2],M=M_{V}\left[1-\frac{6a\sqrt{\gamma}R^{2}}{R^{3}_{\rm SM}f\left(R/R_{\rm SM}\right)\left(1-R^{2}/R^{2}_{\rm SM}\right)^{1/2}}\right], (46)

where γ\gamma is the dimensionless constant of the gradient correction to the kinetic energy density; see Eq. (28).

Refer to caption

Figure 2: NS masses M(R)M(R) [Eq. (45) in Solar units MM_{\odot}] as function of the ES curvature radius RR (in km) are shown by black ρ¯/ρ0=1\overline{\rho}/\rho_{0}=1, red 2, green 3 and blue 4 solid lines, where ρ0=0.16\rho_{0}=0.16 fm-3, for a leptodermic parameter a/R=0.08a/R=0.08. Similar dashed lines are the corresponding volume masses MVM_{V}, Eq. (40). Sensitivity of mass MM for the a/Ra/R variation is shown by the comparison of the green solid (a/R=0.08a/R=0.08) and dash-dotted (0.060.06) lines. The behavior of the gravitational coefficient ff is seen from comparison of the green solid versus the dotted (f=1f=1 at λ=0\lambda=0) curve (both at the same a/R=0.08a/R=0.08). Red, orange, and green spots show the experimental data on the NS J0030+0441 (Ref. 67), GW170817 (Ref. 68), and J0740+6620 (Ref.69), respectively.

Refer to caption   Refer to caption

Figure 3: Contour plots for the NS mass M(R)M(R) (Eq. (45) in Solar units MM_{\odot}) as function of the ES curvature radius RR (in km), and central density ρc\rho_{c} (in nuclear saturation-density units ρ0=0.16\rho_{0}=0.16 fm-3) for a leptodermic parameter a/R=0.08a/R=0.08 (left) and 0.060.06 (right). Red areas show non-physical regions R>RSMR>R_{\rm SM}, and white ones present indeterminations near R=RSMR=R_{\rm SM}.

Figure 2 shows the mass distribution M(R)M(R), Eq. (45), as function of the effective radius RR by using only two physical parameters, the relative crust thickness a/Ra/R, and the central NS particle-number density ρ¯\overline{\rho} in units of the nuclear matter-saturation density ρ0=0.16\rho_{0}=0.16 fm-3, ρ¯/ρ0=14\overline{\rho}/\rho_{0}=1-4, as examples. The surface effect is measured by the relative difference between the full NS mass, M=MV+MSM=M_{V}+M_{S} [see Eq. (45) with Eqs. (40) and (41), solid lines] and its corresponding volume component MVM_{V} (a=0a=0, dashed curves). As seen from Fig. 2, the surface component is rather notable even at a small leptodermic parameter a/R=0.08a/R=0.08. The significant influence of the gravitational forces through the Schwarzschild metric (36) is shown by comparing the solid [Eq. (41) for ff] and dotted (f=1f=1) green lines, both taken at a/R=0.08a/R=0.08. It depends also essentially on the value of a/Ra/R through the equilibrium condition; cf. dashed-dotted (a=0.06a=0.06) and solid (a=0.08a=0.08) green lines. The NS mass M(R)M(R) is the non-monotonic function of RR for a constant central NS density, ρ¯\overline{\rho}, because of the surface component, MSM_{S}, Eq. (44). This is in contrast to the monotonic behavior of the volume mass, MVR3M_{V}\propto R^{3}, in the Cartesian case of the small Newtonian gravitational limit at a=0a=0; see also the dashed (a=0a=0) at finite f(R/RSM)f(R/R_{\rm SM}) [Eq. (41)] curves. For any given value of ρ¯\overline{\rho}, one finds a rather pronounced maximum in dependence of the full mass M(R)M(R), Eq. (45). As well known[2, 3], the physical values of the NS radius for the Schwarzschild metric, Eq. (36) inside of the system, have to obey the condition, rg<R<RSMr_{g}<R<R_{\rm SM}. This means on the left side of the sharp addiction decline, i.e. more pronounceable near the maxima in Fig. 2. These maxima are increased from about 1.2 to 2.5 MM_{\odot} for decreasing ρ¯/ρ0\overline{\rho}/\rho_{0} from 4 to 1 at a/R=0.08a/R=0.08 as example, respectively. The NS mass for each of these curves at a given central value, ρc=ρ¯\rho_{c}=\overline{\rho}, disappears sharply in the limit RRSMR\rightarrow R_{\rm SM}, and does not exist at R>RSMR>R_{\rm SM}. As mentioned above, we should emphasize that our derivations for the surface component of the NS mass, MSM_{\rm S}, fail near the point R=RSMR=R_{\rm SM} because of the obvious singularity of the Jacobian J(r)J(r), Eq. (38), at r=R=RSMr=R=R_{\rm SM}. As seen from Fig. 2, our results are in reasonable agreement with the experimental data[68, 67, 69]; see also Ref. 70 for the theoretical results about the neutron stars. For smaller NS masses, M(1.21.5)MM\approx(1.2-1.5)M_{\odot} for the NS pulsars GW170817 (Ref. 68) and J0030+0451 (Ref. 67) with the slightly different radii, R=10.311.8R=10.3-11.8 km (orange spot) and R=11.613.1R=11.6-13.1 km (red spot), one finds respectively good agreement with our results for the central density ρ¯=3ρ0\overline{\rho}=3\rho_{0}, with a/R=0.08a/R=0.08. Larger mass of the NS pulsar J0740+6620 (Ref. 69), M=(2.02.1)MM=(2.0-2.1)M_{\odot} and radius R=11.313.6R=11.3-13.6 km (green spot), corresponds to the same central density, ρc=ρ¯\rho_{c}=\overline{\rho}, but with smaller leptodermic parameter a/R=0.06a/R=0.06.

Figure 3 shows contour plots for the NS mass, MM (in Solar units) as functions of its radius, RR, and central density, ρc\rho_{c}, (ρ¯\overline{\rho} in units of ρ0=0.16\rho_{0}=0.16fm-3) for the leptodermic parameter values, a/R=0.08a/R=0.08 (left) and 0.060.06 (right). As seen from the left figure, one for instance finds the line for the NS mass 1.5M1.5M_{\odot} for radius R=12.614.0R=12.6-14.0 km and central density, ρ¯/ρ0=1.82.7\overline{\rho}/\rho_{0}=1.8-2.7. In the right figure we can see the line for the NS mass 2.0M2.0M_{\odot} for radius R=10.914.0R=10.9-14.0 km and central density, ρ¯/ρ0=1.84.3\overline{\rho}/\rho_{0}=1.8-4.3. They are close to the experimental data shown in Fig. 2. The NS mass structure is changed significantly with radius RR and relative central density ρc\rho_{c}, also with the leptodermic parameter. The calculations become unstable with approaching the line R=RSMR=R_{\rm SM}, as shown by white areas. The physical region is below this line because of the restriction by the condition rg<R<RSMr_{g}<R<R_{\rm SM}.

6 The NS surface energy

Let us derive now the NS total energy EE which is similar to the main two (volume and surface) terms of the Weizsäker mass formula, basic in nuclear physics, but with the analytical expression for its tension coefficient and accounting for the gravitational forces. According to the basic definitions, the total energy EE, Eq. (2), can be presented as

E=d𝒱[ρ(𝐫)]EV+ES,E=\int\hbox{d}\mathcal{V}\mathcal{E}[\rho({\bf r})]\approx E_{V}+E_{S}~{}, (47)

where d𝒱\hbox{d}\mathcal{V} is given by Eqs. (35) and (38), and EVE_{V} is the volume part of the total energy. Following the mass derivations in the previous section, one has

EV=bV(G)ρ¯d𝒱=4πbV(G)ρ¯RSM3f(RRSM),E_{V}=-b^{(G)}_{V}\overline{\rho}\int\hbox{d}\mathcal{V}=-4\pi b^{(G)}_{V}\overline{\rho}R_{\rm SM}^{3}f\left(\frac{R}{R_{\rm SM}}\right), (48)

where bV(G)b^{(G)}_{V} is given by Eq. (9), and f(z)f(z) is shown in Eq. (41). In Eq. (47), ESE_{S} is the surface part,

ESσS,E_{S}\approx\sigma S, (49)

where S=4πR2S=4\pi R^{2} is the surface area value, and σ\sigma is the tension coefficient. The leading expression for σ\sigma in our leptodermic expansion over small parameter a/Ra/R is given by444 Similar expressions as Eqs. (50) and (51) at 𝒟=Γ=0\mathcal{D}=\Gamma=0 were obtained early by van der Waals; see Ref. 36.

σ=2J(R)dξ[𝒞+𝒟ρ+Γ4ρ](ρξ)2,\sigma=2J(R)\int_{-\infty}^{\infty}\hbox{d}\xi\left[\mathcal{C}+\mathcal{D}\rho+\frac{\Gamma}{4\rho}\right]\left(\frac{\partial\rho}{\partial\xi}\right)^{2}, (50)

where J(R)J(R) is the Schwarzschild metric Jacobian J(r)J(r) at r=Rr=R; see Eq. (38). In these derivations, we present locally the integration over d𝒱\hbox{d}\mathcal{V} as that over the system surface and the normal coordinate dξ\hbox{d}\xi, d𝒱=Jgdηdξ\hbox{d}\mathcal{V}=J\sqrt{g}\hbox{d}\eta\hbox{d}\xi, where g\sqrt{g} is the Newtonian volume element related to the metrics given in Eq. (77) [see Eq. (76) for g\sqrt{g}] and Eq. (38) for the Jacobian J(r)J(r). For the spherical system, one has g=1\sqrt{g}=1. Then, we expand the integration limits of ξ\xi to ±\pm\infty because (ρ/ξ)2(\partial\rho/\partial\xi)^{2} has a sharp maximum at the ES (ξ=0\xi=0). Factor 2 appears because of Eq. (24) which leads to the approximate (at leading order in small parameter a/Ra/R) equivalence of the second order [εG(ρ)\varepsilon_{G}(\rho)] correction to the bulk energy density bV(G)ρ¯-b^{(G)}_{V}\overline{\rho}, and the surface (gradient) parts of the total energy EE. As we have square of density derivative over ξ\xi in Eq. (50), one can use Eq. (24) for the density ρ\rho at the leading order. Equation (50) can be transformed to the expression with the dimensionless integral over yy, that is more convenient for calculations,

σ=aρ¯KG9J(R)01dy(1+βy+γ/y)dydxaρ¯KG9J(R)01dy(1+βy+γ/y)ϵ(y),\sigma=-\frac{a\overline{\rho}K_{G}}{9}J(R)\!\int_{0}^{1}\!\hbox{d}y(1+\beta y+\gamma/y)\frac{{\rm d}y}{{\rm d}x}\approx\frac{a\overline{\rho}K_{G}}{9}J(R)\!\int_{0}^{1}\!\hbox{d}y\sqrt{\left(1+\beta y+\gamma/y\right)\epsilon(y)}, (51)

where J(R)J(R) is given by Eq. (38) at r=Rr=R. In these derivations we transform the variable ξ\xi to ρ\rho, dξ=(dξ/dρ)dρ\hbox{d}\xi=(\hbox{d}\xi/\hbox{d}\rho)\hbox{d}\rho, as above, and use Eqs. (23), (24), and (26)-(28). For the approximation (19) to ϵ(y)\epsilon(y), the integral in Eq. (51) can be taken in terms of the elementary functions. The simplest analytical expression for the tension coefficient σ\sigma is obtained for β=γ=0\beta=\gamma=0 [case (i)]. Using Eq. (32) for y=y(x)y=y(x), one finally arrives at

σ=4aρ¯KG135J(R).\sigma=4a\frac{\overline{\rho}K_{G}}{135}J(R)~{}. (52)

Similarly, the expression for σ\sigma in the Wilets case (ii) can be derived too, σ=aρ¯KGJ(R)γ/36\sigma=a\overline{\rho}K_{G}J(R)\sqrt{\gamma}/36. Thus, one obtains the explicitly analytical expressions for the tension coefficient σ\sigma as functions of constants of the energy density \mathcal{E}, Eq. (3), and Schwarzschild metric Jacobian at the ES, J(R)J(R). For nuclear physics, the tension coefficient σ\sigma is related to the vdW-Skyrme interaction constants, in good agreement with the van der Waals capillary theory[36].

Equation (2) is the two main (volume and surface) terms of the usual leptodermic expression for the total energy of the dense liquid drop. In nuclear physics, this expression was obtained with including additional terms of the symmetry energy [with volume (and Swiatecki linear term) and surface (isovector) corrections] and curvature and Coulomb components (see, for instance, Refs. 34, 35 for the ESA).

Notice that due to the integration over the normal-to-ES radial variable ξ=rR\xi=r-R, the smallness factor, proportional to a/Ra/R, appears for the surface with respect to the volume contributions to the NS mass, Eq. (44), and total energy EE, Eq. (47). Thus, these surface components, MSM_{S} and ESE_{S} are of the order of a small parameter a/Ra/R. However, these surface contributions, MSM_{S} and ESE_{S}, dramatically influences on the total NS characteristics because the vdW capillary surface pressure (including the gravitational forces) equilibrates the volume pressure acting from the interior of the liquid-drop to be a leading reason of its equilibrium stability. We will study this stability condition in more details in the next section.

Refer to caption   Refer to caption


Figure 4: Pressure PP, Eq. (60), in units of ρ¯KG/9\overline{\rho}K_{G}/9 as function of the density variable y=ρ/ρ¯y=\rho/\overline{\rho} in (a) for the universal EoS, and of the radial coordinate rr in (b) through the particle number density ρ=ρ¯y((rR)/a)\rho=\overline{\rho}y((r-R)/a), where y(x)y(x) is given, e.g., by Eq. (32) (solid lines). Dashed (red) and dotted (green) lines show the surface (SS) and volume (VV) components. The crust thickness a=1a=1 km and the effective radius parameter R=10R=10 km are the same as in Fig. 1(b). The full dots and arrows present the ES for y=y0=1/3y=y_{0}=1/3 (a) [see Eq. (32)] and r=Rr=R (b).

7 Equation of state

The equation of state (EoS), P=P(ρ)P=P(\rho), can be determined[71] by using the energy density (ρ)\mathcal{E}(\rho), Eq. (3),

P=ρ2δ𝒲δρ.P=\rho^{2}\frac{\delta\mathcal{W}}{\delta\rho}~{}. (53)

The energy per particle 𝒲\mathcal{W} is a function of the particle number density ρ\rho; see Eqs. (5) for 𝒲\mathcal{W}, and (3) for the energy density \mathcal{E}. According to Eq. (14) for the equilibrium of the dense liquid-drop system, one obtains the pressure PP in the form:

P=μρ(ρ)=μρ𝒜(ρ)(ρ)(ρ)2,P=\mu\rho-\mathcal{E}(\rho)=\mu\rho-\mathcal{A}(\rho)-\mathcal{B}(\rho)(\nabla\rho)^{2}~{}, (54)

where μ\mu is the chemical potential, and (ρ)\mathcal{E}(\rho) is the energy density, Eq. (3). The leading term of the pressure (54) can be presented as

PρεG(ρ)(ρ)(ρξ)2,P\approx\mathcal{M}\rho-\varepsilon_{G}(\rho)-\mathcal{B}(\rho)\left(\frac{\partial\rho}{\partial\xi}\right)^{2}~{}, (55)

where =μ+bV(G)\mathcal{M}=\mu+b^{(G)}_{V}; see Eq. (21), ρ/ξ\partial\rho/\partial\xi is given by Eq. (24) in terms of the εG(ρ)\varepsilon_{G}(\rho), Eq. (23), (ρ/ξ)2=εG(ρ)\mathcal{B}(\partial\rho/\partial\xi)^{2}=\varepsilon_{G}(\rho). For a leptodermic system, a/R1a/R\ll 1, one can rewrite this equation as

PPV+PS,P\approx P_{V}+P_{S}~{}, (56)

where PVP_{V} is the volume part of the pressure PP [see Eq. (16)]

PVμρ𝒜(ρ)=ρεG(ρ).P_{V}\equiv\mu\rho-\mathcal{A}(\rho)=\mathcal{M}\rho-\varepsilon_{G}(\rho)~{}. (57)

The second component, PSP_{\rm S}, is the surface part,

PS(ρ)(ρξ)2εG(ρ)ρ¯KG18y(1y)2.P_{S}\equiv-\mathcal{B}(\rho)\left(\frac{\partial\rho}{\partial\xi}\right)^{2}\approx-\varepsilon_{G}(\rho)\approx-\frac{\overline{\rho}K_{G}}{18}y(1-y)^{2}. (58)

For the middle we used Eq. (24) for the particle number density ρ\rho in the leading order approximation. Using Eqs. (23) and (19) for the quadratic approximation, one arrives at the last equation. Notice that the surface component, PSP_{S}, is concentrated near the effective surface, r=Rr=R, with the exponential decrease inside and outside of the mass system; see Fig. 4. According to Eqs. (57) and (58) for the quadratic approximation to ϵ(y)/y\epsilon(y)/y, Eq. (19), one has a macroscopic liquid-drop equilibrium condition at the ES, (PV+PS)r=R=0(P_{V}+P_{S})_{r=R}=0. Thus, the chemical potential surface component \mathcal{M} at the liquid-drop equilibrium equals

=KG9(y01)2,\mathcal{M}=\frac{K_{G}}{9}(y_{0}-1)^{2}, (59)

where y0y_{0} is the value of the density y(x)y(x) at the ES, x=0x=0, y0=1/3y_{0}=1/3 for the vdW-Skyrme case (i), and y0=1/2y_{0}=1/2 for the Wilets solution (ii). With Eq. (59), one finally obtains from Eqs. (56), (57), and (58),

PKG9y[(y01)2(y1)2]=PV+PS,P\approx\frac{K_{G}}{9}~{}y\left[(y_{0}-1)^{2}-(y-1)^{2}\right]=P_{\rm V}+P_{\rm S}~{}, (60)

where

PV=KG9y[(y01)212(y1)2)],\displaystyle P_{V}=\frac{K_{G}}{9}~{}y\left[(y_{0}-1)^{2}-\frac{1}{2}(y-1)^{2})\right]~{}, (61)
PS=KG18y(y1)2.\displaystyle P_{S}=-\frac{K_{G}}{18}~{}y(y-1)^{2}~{}. (62)

Figure 4 shows the pressure PP as function of the dimensionless density yy, (a), and radial variable rr, (b), through the density, y((rR)/a)y((r-R)/a), at leading order over the parameter a/Ra/R. See Eq. (60) for the full pressure, Eq. (61) for the volume (PVP_{V}), and Eq. (62) for the surface (PSP_{\rm S}) components of the pressure. As seen from Fig. 4(a), all pressures, PP, PVP_{V} and PSP_{S}, take zero value at y=0y=0. The surface pressure, PSP_{S}, is zero also asymptotically in the limit y1y\rightarrow 1, in contrast to the volume part, PVP_{V} (the total pressure PP). The full circle shows the ES at y=y0=1/3y=y_{0}=1/3 for the (i) case. As expected, the volume pressure PVP_{V} is monotonically increasing function of the density yy. In Fig. 4(b), the pressure P(ρ(r))P(\rho(r)) (solid) has a typical leptodermic behavior with a relatively short thickness of the order of aa, where the pressure decreases sharply from almost the constant, which is the asymptote inside of the system at r > Rar\hbox{\kern 1.00006pt\lower 2.58334pt\hbox{$\sim$} \kern-11.19997pt\raise 2.58334pt\hbox{$>$} \kern 1.00006pt}R-a, to zero. Dashed red and dotted green lines show the surface (PSP_{S}) and volume (PVP_{V}) components, respectively. Sum of these components is the total pressure PP with zero at the ES; see the full point (ES) and arrow which present the effective surface and radius RR, respectively. As seen also from Fig. 4(b), we obtain analytically a sharp minimum of the surface pressure, PS(r)P_{\rm S}(r), near the effective surface (rRr\approx R) with its exponential decrease inside and outside of the ES to zero. In the vdW-Skyrme case (i), the pressure P(r)P(r) is exponentially sharp decreasing function of rr outside of the system. Notice that, in contrast to this, the Wilets solution in the symmetric case (ii) for the particle density ρ\rho, Eq. (33), leads to a large tail outside of the system as expected from Fig. 1(b). For the parameters R=10R=10 km, and a=1a=1 km (see Figs. 4 and 1(b)), one finds that the pressure P(r)P(r) [Eq. (60)] disappears in the vdW-Skyrme case (i) at r > R+ax0r\hbox{\kern 1.00006pt\lower 2.58334pt\hbox{$\sim$} \kern-11.19997pt\raise 2.58334pt\hbox{$>$} \kern 1.00006pt}R+ax_{0} (x01.32x_{0}\approx 1.32), that corresponds to about r > 11.3r\hbox{\kern 1.00006pt\lower 2.58334pt\hbox{$\sim$} \kern-11.19997pt\raise 2.58334pt\hbox{$>$} \kern 1.00006pt}11.3 km, similarly as for the density ρ(r)\rho(r), which is a very close distance from the ES. This distance is decreased much for the pressure P(r)P(r) in the case of the significantly small edge thickness a=0.1a=0.1 km to 10.7 km, that is somehow larger than for the particle number density.

8 Tolman-Oppenheimer-Volkoff approach

The basic relations determining the mass-radius relation are the traditional TOV differential equations[1, 4, 2, 18]:

dPdr=G(+P)(mc2+4πr3P)rc2(rc22Gm),dmdr=4πr2c2,\frac{{\rm d}P}{{\rm d}r}=-\frac{G(\mathcal{E}+P)(mc^{2}+4\pi r^{3}P)}{rc^{2}(rc^{2}-2Gm)},\quad\frac{{\rm d}m}{{\rm d}r}=\frac{4\pi r^{2}\mathcal{E}}{c^{2}}~{}, (63)

where m(r)m(r) is the gravitational mass interior to the radius rr, \mathcal{E} is the energy density, Eq. (3) in our approach. Boundary conditions for these equations are

m(r=0)=0,(dP/dr)r=0=0.m(r=0)=0,\quad(\hbox{d}P/\hbox{d}r)_{r=0}=0~{}~{}. (64)

The integrations are terminated when P=0P=0, which defines the surface r=Rr=R. A specified value of the central pressure P0=P(r=0)P_{0}=P(r=0) determines the total mass M=m(r=R)M=m(r=R) in the TOV approach. Note that the only EoS relation needed is the pressure density relation P=P(ρ)P=P(\rho).

Taking approximately into account the step-function density, ρ=ρ¯\rho=\overline{\rho} for r<Rr<R inside of the system, and ρ=0\rho=0 for r>Rr>R outside of it, one can solve TOV equations (63) analytically. Our leading-order solution, Eq. (27), and examples, Eqs. (32) and (33) [see Fig. 1(b)], having a sharp transition from the saturation to zero value, largely agree with above mentioned approximation at small leptodermic parameter a/R1a/R\ll 1, the better the smaller this parameter. The step-like density can be considered as the zero order in the leptodermic expansion in the limit a/R0a/R\rightarrow 0. Taking into account the first boundary condition in Eq. (64), the second equation in Eq. (63) can be integrated analytically,

m(r)=4π03c2r3,m(r)=\frac{4\pi\mathcal{E}_{0}}{3c^{2}}r^{3}, (65)

where

0=(ρ=ρ¯).\mathcal{E}_{0}=\mathcal{E}(\rho=\overline{\rho}). (66)

This value should be some constant, independent of rr for the assumed coordinate dependence of the density ρ(r)\rho(r) inside of the particle system. In particular, it can be taken approximately as shown in Eqs. (3), (9), and (23). Therefore, the total mass MM in this approach is given by

M=m(R)=4π03c2R3,M=m(R)=\frac{4\pi\mathcal{E}_{0}}{3c^{2}}R^{3}, (67)

where RR is the effective NS radius (r1r_{1} in the notations of Tolman’s book, Ref. 2). Notice that the NS mass MM, Eq. (44), derived in Section 5, differs essentially from this expression, Eq. (67), by the surface component MSM_{\rm S}; see Eq. (44). Another essential difference comes from the Schwarzschild metric Jacobian JJ, Eq. (38), with the influence on the surface and volume parts of the NS mass.

Substituting Eq. (65) into the first TOV Eq. (63), one can see that the variables PP and rr are separated. Therefore, we analytically find the solution for rRRSMr\leq R\leq R_{\rm SM} and (P+0/3)/(P+0)>0(P+\mathcal{E}_{0}/3)/(P+\mathcal{E}_{0})>0:

P=033ζ1r2/RSM211ζ1r2/RSM2,P=\frac{\mathcal{E}_{0}}{3}\frac{3\zeta\sqrt{1-r^{2}/R_{\rm SM}^{2}}-1}{1-\zeta\sqrt{1-r^{2}/R_{\rm SM}^{2}}}, (68)

where

ζ=|P0+0/3P0+0|,\zeta=\Big{|}\frac{P_{0}+\mathcal{E}_{0}/3}{P_{0}+\mathcal{E}_{0}}\Big{|}~{}, (69)

and P0P(ρ¯)=P(r=0)P_{0}\equiv P(\overline{\rho})=P(r=0). In Eq. (68), RSMR_{\rm SM} is the radius parameter of the Schwarzschild metric in the form presented in Eq. (36); see Eq. (37) for RSMR_{\rm SM}, and Ref. 2. This metric is non-singular at r=0r=0. It was derived from the original form, valid outside of the NS matter system, Refs. 2, 3, and singular at r=0r=0, by using a transformation of coordinates and GRT invariance; see, e.g., Ref. 2.

Refer to caption

Figure 5: Contour plots for the pressure P(r)P(r), Eq. (70), in units of the central value, P(r=0)P(r=0), as function of the radial coordinate rr and the dimensionless parameter ASMA_{\rm SM} of the Schwarzschild metric, Eq. (36). The numbers in squares show the values of this pressure. White color presents regions where we have indetermination, infinity by infinity, with the finite limit 1 at r0r\rightarrow 0. Red color shows negative values of the ratio P(r)/P(r=0)P(r)/P(r=0) [a positive pressure, P(r)P(r)]. The value “0.01” displays approximately the zero value in horizontal and vertical lines on right of plots. The effective NS radius R=10R=10 km is the same as in Figs.  4 and 1(b).

Refer to caption

Figure 6: The pressure P(r)P(r), Eq. (70), in units of the central value, P(r=0)P(r=0) as function of the radial coordinate rr for the Schwarzschild metric in the form Eq. (36) at two values of the parameter ASMA_{\rm SM}, ASM=1.15A_{\rm SM}=1.15 (solid line) and ASM0.96A_{\rm SM}\approx 0.96 (dashed line) which correspond approximately to the two limit values for the NS masses, M=1.4MM=1.4M_{\odot} and M=2.0MM=2.0M_{\odot}, which are related to the recent experimental data[67, 68, 69] (RSM13.0R_{\rm SM}\approx 13.0 km and 15.615.6 km, respectively ; see text). The effective NS radius R=10R=10 km is the same as in Fig.  4.

Notice that the solution (68) coincides, up to a constant related to units, with that presented in Tolman’s book (Ref. 2) at ζ=BSM/ASM=1/(2ASM)\zeta=B_{\rm SM}/A_{\rm SM}=1/(2A_{\rm SM}),

P=033BSM1r2/RSM2ASMASMBSM1r2/RSM2,P=\frac{\mathcal{E}_{0}}{3}~{}\frac{3B_{\rm SM}\sqrt{1-r^{2}/R^{2}_{\rm SM}}-A_{\rm SM}}{A_{\rm SM}-B_{\rm SM}\sqrt{1-r^{2}/R^{2}_{\rm SM}}}~{}, (70)

where

BSM=12,ASM=12ζ.B_{\rm SM}=\frac{1}{2},\quad A_{\rm SM}=\frac{1}{2\zeta}~{}. (71)

We have to choose the sign plus on right of last equation for the positive expression of (P+0/3)/(P+0)(P+\mathcal{\mathcal{E}}_{0}/3)/(P+\mathcal{\mathcal{E}}_{0}), because the minus, which is related to the opposite sign of the same expression, is not valid for the Schwarzschild metric, ASM>0A_{\rm SM}>0. The cosmological constant is assumed to be zero, 0RSM2\mathcal{E}_{0}\propto R^{-2}_{\rm SM}, Eq. (37), (G=c=1G=c=1 in Tolman’s book). The constants, ASMA_{\rm SM}, BSMB_{\rm SM}, and RSMR_{\rm SM} are parameters of the transformed interior Schwarzschild metric, Eq.  (36), for ASM>0A_{\rm SM}>0,

ASM=321R2RSM2,BSM=12.A_{\rm SM}=\frac{3}{2}\sqrt{1-\frac{R^{2}}{R_{\rm SM}^{2}}},\quad B_{\rm SM}=\frac{1}{2}~{}. (72)

Therefore, one obtains the relation of the parameters of the Schwarzschild metric, Eq.  (36), to the initial condition P=P0P=P_{0} at r=0r=0 through Eq. (69). Notice that according to Eq. (70) for the pressure P(r)P(r), one has P=0P=0 at the boundary r=Rr=R because of Eq. (72) for the constants ASMA_{\rm SM} and BSMB_{\rm SM}. Finally, one has the full agreement of our result for the pressure, Eq. (68), with that of Ref. 2.

9 Discussions of the results

Figure 5 shows the contour plots for the pressure P(r,ASM)P(r,A_{\rm SM}) [in units of P(r=0,ASMP(r=0,A_{\rm SM})], as function of rr (km) and parameter ASMA_{\rm SM} of the Schwarzschild metric (36); see Eq. (72). It is more convenient to use the finite dimensionless values of ASMA_{\rm SM}, 0ASM1.50\leq A_{\rm SM}\leq 1.5, instead of dimensional RSMR_{\rm SM} in the infinite range for the fixed effective NS surface radius R=10R=10 km, 0rR0\leq r\leq R. For a rough estimate 0c2ρc\mathcal{E}_{0}\sim c^{2}\rho_{c}, where the central mass density ρcM/(4πR3/3)\rho_{c}\approx M/(4\pi R^{3}/3) (MMVM\approx M_{V}), and M/M1.42.0M/M_{\odot}\approx 1.4-2.0 in units of the Solar mass MM_{\odot}, one finds significantly larger parameter RSMR_{\rm SM} than the effective radius RR, RSM15.613.0R_{\rm SM}\approx 15.6-13.0 km in this case555 Notice that the Schwarzschild parameter RSMR_{\rm SM} differs essentially from the gravitational radius rgr_{\rm g}, introduced by Schwarzschild, rg=2GM/c2Rr_{\rm g}=2GM/c^{2}\leq R; see Ref. 3. Indeed, RSMR_{\rm SM} is assumed to be larger than the effective NS radius RR., respectively. With these evaluations, one has respectively the constants ASM1.150.96A_{\rm SM}\approx 1.15-0.96 for M(1.42.0)MM\approx(1.4-2.0)M_{\odot}. Equation (71) presents the definite relation between this ASMA_{\rm SM} and the dimensionless variable ζ\zeta, Eq. (69), which can be also considered equivalently instead of the two dimensional values of the initial pressure P0P_{0} and energy density 0\mathcal{E}_{0}. For the gravitational radius, one obtains rg=4.15.9r_{\rm g}=4.1-5.9 km, that is smaller than the effective radius R=10R=10 km. The maximum value of ASMA_{\rm SM}, ASM=1.5A_{\rm SM}=1.5, corresponds to zero pressure P(r)P(r).

Figure 6 shows the two cuts of the contour plots (Fig. 5) at the value ASM=1.15A_{\rm SM}=1.15 (RSM15.6R_{\rm SM}\approx 15.6 km, solid black line), and the value ASM=0.96A_{\rm SM}=0.96 (RSM13.0R_{\rm SM}\approx 13.0 km, dashed red line). The exemplary physical region, M/M1.42.0M/M_{\odot}\approx 1.4-2.0, which are consistent with the recent experimental data shown in Fig. 2 for NS masses (see Refs. 67, 68, 69, also Refs. 15, 20, 6, 70), is related to a small range in Fig. 6. These curves have the same starting ordinate 1 and final one 0 .

For more realistic values of RSMR_{\rm SM}, for instance, taking into account the gravitational defect of mass[3] in the evaluations of the central energy density 0\mathcal{E}_{0}, one may have different values of ASMA_{\rm SM} presented in Fig. 5. Thus, as seen from solid and dashed lines of Fig. 6 and those assumed in the derivations of the TOV equations[4, 2], the pressure PP turns, indeed, to zero at the boundary of the NS system, r=Rr=R.

Refer to caption

Figure 7: Left (a): The pressure versus r of the NS with the Skyrme interaction KDE0v1 by solving TOV equations numerically. Right (b): The pressure versus the corresponding density of the NS. The central density ρc=0.64\rho_{c}=0.64 fm-3 and the mass of the NS is 1.62 M.

Refer to caption   Refer to caption


Figure 8: Left: The density profiles of the NS with the Skyrme interaction KDE0v1 by solving TOV equations numerically. The black solid line refers to the central density ρc=0.64\rho_{c}=0.64 fm-3 and the mass of the NS is 1.62 M. The red dotted line refers to the central density ρc=0.64\rho_{c}=0.64 fm-3 with changing the gravitational constant G to G/4 and the mass of the NS is 4.0 M. The blue dashed-dotted line refers to another central density ρc=0.416\rho_{c}=0.416 fm-3 and the mass of the NS is 1.02 M. Right: The density profiles of the NS with the Skyrme interaction KDE0v1 by solving TOV equations numerically. The black solid line refers to the central density ρc=0.16\rho_{c}=0.16 fm-3 with changing the gravitational constant G to G/4 and keeping the mass of the NS to be 1.62 M. The red dotted line is the WS fit, Eq. (73); see the text.

Figure 7 shows the results obtained numerically by solving the TOV equations (63) for the pressure P=P(r)P=P(r) (a) with the EoS, P=P(ρ)P=P(\rho) (b), related to the Skyrme force KDE0v1 [72]; see Refs. 72, 15, 73. For convenience, the radial coordinate rr in (a) is presented in km units while the density variable ρ\rho is done in nuclear units fm-3. In both panels (a) and (b), the same pressure PP is displayed in nuclear units MeV/fm3. The pressure P(r)P(r) decreases with increasing rr while the EoS function of ρ\rho, P=P(ρ)P=P(\rho), has an opposite behavior. These numerical semi-microscopic results are qualitatively in agreement with our macroscopic volume components PVP_{V} shown in Fig. 4 inside of the NS.

The particle number density ρ\rho is shown in Fig. 8(a) as function of the radial coordinate rr (km) for different parameters. The masses M/M=1.62M/M_{\odot}=1.62 and 1.021.02, larger than the Solar mass MM_{\odot}, correspond to different values of the central density ρc=0.64\rho_{c}=0.64 fm-3 (solid black) and 0.4160.416 fm-3 (blue dashed-dotted curve), significantly larger than the nuclear matter value 0.160.16 fm-3. All of them are related to the EoS KDE0v1 Skyrme force[72]; see also Refs. 15, 73. The gravitational force effects are tested formally for decreasing the gravitational constant GG by factor 4 in red dotted line in Fig. 8(a). The Woods-Saxon leptodermic fit is shown by the red dotted line in Fig. 8(b). The last situation is qualitatively close to our analytical results for the volume (V) contributions presented in Figs. 4-6. The main difference in Fig. 4(b), as compared to the numerical results in Figs. 7,8, is the surface contribution (S) into EoS, P=P(y)P=P(y) (y=ρ/ρ¯y=\rho/\overline{\rho}) [see Fig. 4 (a)] and the pressure leptodermic correction PS(r)P_{\rm S}(r), Eq. (62) [Fig. 4(b)]. The shape of the total pressure behavior, P=P(r)P=P(r), is more leptodermic in Fig. 4(b) than in that of Fig. 7 (left). These results are in good agreement with the density shapes for the analytical results in Fig. 1(b) and numerical calculations in Fig. 8 (right) as displayed by the red curve for the inversed Woods-Saxon (WS) fit. The WS fit formula is given by

ρ(r)=ρWS1+exp[(rRWS)/aWS],\rho(r)=\frac{\rho_{\rm WS}}{1+\hbox{exp}\left[\left(r-R_{\rm WS}\right)/a_{\rm WS}\right]}~{}, (73)

where ρWS\rho_{\rm WS}=0.16 fm-3, RWS=14.2786R_{\rm WS}=14.2786 km, and aWS=0.7074a_{\rm WS}=0.7074 km. The results in Figs. 7(a) and 8 (right, red dotted line) for the volume contributions of the pressure P(r)P(r), obtained analytically from the TOV equations (63), are naturally similar to those shown in Fig. 6, and Fig. 4 (b) for the volume contribution. The surface contributions to the standard TOV equations (63) and the corresponding solutions will be studied in the forthcoming work where we are going to take into account explicitly the ES corrections.

10 Conclusions

The effective surface approximation based on the leptodermic expansion over a small parameter - a ratio of the crust thickness aa to the effective radius RR of the system - is extended for the description of neutron star properties. Neutron star was considered as a dense liquid drop at the equilibrium. The gravitational potential Φ(ρ)\Phi(\rho) was taken into account in the simplest form as expansion over powers of differences ρρ¯\rho-\overline{\rho}, where ρ¯\overline{\rho} is the saturation density, up to second order in terms of the separation particle energy and incompressibility. For a strong gravitation, one has its significant contribution to the separation particle energy bV(G)b^{(G)}_{V} and incompressibility modulus KGK_{G}, within the Schwarzschild metric solution to the General Relativity Theory equations. Taking into account the gradient terms of the energy density in a rather general form we analytically obtained the leading (over a small parameter a/R1a/R\ll 1) particle number density ρ\rho as function of the normal-to-ES coordinate ξ\xi of the orthogonal local nonlinear-coordinate system ξ,η,φ\xi,\eta,\varphi through the effective surface. This result is in a good agreement with the van der Waals phenomenological capillary theory. With the help of the local coordinate system ξ,η\xi,\eta, one finds the equation of state (EoS) for the pressure P=P(ρ)P=P(\rho), and the NS surface mass and energy corrections for a spherical NS in the leading ES approach. Our results for the dependence of the NS mass MM on the NS radius RR, including the gravitational effects through the Schwarzschild metric, are in a reasonable agreement with their recent experimental data for several neutron stars. Adding a first order correction over a/Ra/R, one obtains also the NS energy as a sum of the volume and surface terms with the analytical expression for the surface tension coefficient, also in agreement with the van der Waals theory. We obtained the effective surface corrections to the pressure PP of the EoS, which are significant near the ES for macroscopic condition of the NS stability. In line of Tolman derivations[2], for the step-like particle number density, the TOV equations were solved analytically in terms of the initial values of the pressure and energy density at zero radial coordinate. The volume contributions to the pressure in our macroscopic analytical calculations are in good agreement with the semi-microscopic numerical results obtained from the TOV equations with the EoS based on the Skyrme forces, in the case of leptodermic fit of the particle number density ρ\rho to its behavior for a Woods-Saxon potential. From comparison of the analytical and numerical results for the pressure and density, one can conclude importance of the surface corrections near the NS effective surface.

As perspectives, one can generalize our analytical approach to take into account the gradient terms in the TOV equations, also to many-component and rotating systems. It is especially interesting to extend this approach to take into account the symmetry energy of the isotopically asymmetric finite systems. As this approach was formulated in the local nonlinear system of coordinates ξ,η,φ\xi,\eta,\varphi for deformed and superdeformed shapes of the effective surface, one can apply our method to the NS rotating pulsars at large angular momenta.

Acknowledgements

The authors greatly acknowledge C.A. Chin, V.Z. Goldberg, S.N. Fedotkin, J. Holt, C.M. Ko, E.I. Koshchiy, J.B. Natowitz, A.I. Sanzhur, G.V. Rogachev for many creative and useful discussions.

Appendix A Local coordinates near the effective surface

The axially-symmetric shapes of the ES in cylindrical coordinates are assumed to be determined in terms of a certain profile function y=Y(z)y=Y(z) in the plane of axis of symmetry by rotation around zz axis. The ξ\xi is defined in the text as the coordinate perpendicular to the ES. The other coordinate η\eta can be chosen for example, as zz-coordinate of the point at the surface where the perpendicular to the ES from the given point 𝐫{\bf r} crosses it. Thus, let us define new coordinates ξ\xi,η\eta, related to the cylindrical coordinates, as

y=y(ξ,η)=Y(η)+ξ,z=z(ξ,η)=ηξY(η)η,y=y(\xi,\eta)=Y(\eta)+\frac{\xi}{{\cal L}},~{}~{}~{}z=z(\xi,\eta)=\eta-\frac{\xi}{{\cal L}}\frac{\partial Y(\eta)}{\partial\eta}, (74)

where

=[1+(Y(η)η)2]1/2.{\cal L}=\left[1+\left(\frac{\partial Y(\eta)}{\partial\eta}\right)^{2}\right]^{1/2}. (75)

The length and volume elements are given by

dl=(gξdξ2+gηdη2+gφdφ2)1/2,d𝐫=gdξdηdφ,g=(R1+ξ)(1+ξ/R2)\hbox{d}{\it l}=\left(g_{\xi}\hbox{d}\xi^{2}+g_{\eta}\hbox{d}\eta^{2}+g_{\varphi}\hbox{d}\varphi^{2}\right)^{1/2},~{}~{}\hbox{d}{\bf r}=\sqrt{g}\hbox{d}\xi\hbox{d}\eta\hbox{d}\varphi,~{}~{}~{}\sqrt{g}=\left(R_{1}+\xi\right)\left(1+\xi/R_{2}\right) (76)

with the diagonal metric tensor,

gξ=1,gη=(1+ξR2)22,gφ=(1+ξ)2Y2(η),g_{\xi}=1,\qquad g_{\eta}=(1+\frac{\xi}{R_{2}})^{2}{\cal L}^{2},\quad g_{\varphi}=\left(1+\frac{\xi}{{\cal L}}\right)^{2}Y^{2}(\eta), (77)

R1R_{1} and R2R_{2} are two local curvature radii of the ES,

R1=Y(η),R2=32Y/η2.R_{1}={\cal L}Y(\eta),\qquad\qquad R_{2}=-\frac{{\cal L}^{3}}{\partial^{2}Y/\partial\eta^{2}}~{}. (78)

According to the right of Eq. (76), the mean curvature writes

H=12(1R1+1R2)=12(ξlng).H=\frac{1}{2}\left(\frac{1}{R_{1}}+\frac{1}{R_{2}}\right)=\frac{1}{2}\left(\frac{\partial}{\partial\xi}\ln\sqrt{g}\right). (79)

For the Laplace operator in new coordinates ξ\xi and η\eta, one finds

Δξη=2ξ2+ξ(lng)ξ+1gη(ggηη).\Delta_{\xi\eta}=\frac{\partial^{2}}{\partial\xi^{2}}+\frac{\partial}{\partial\xi}\left(\ln\sqrt{g}\right)\;\frac{\partial}{\partial\xi}+\frac{1}{\sqrt{g}}\frac{\partial}{\partial\eta}\left(\frac{\sqrt{g}}{g_{\eta}}\frac{\partial}{\partial\eta}\right). (80)

ORCID
A. G. Magner https://orcid.org/0000-0003-1694-640X
S. P. Maydanyuk https://orcid.org/0000-0001-7798-1271
A. Bonasera https://orcid.org/0000-0001-7147-4535
H. Zheng https://orcid.org/0000-0001-5509-4970
A. I. Levon https://orcid.org/0000-0002-9880-1927
T. M. Depastas https://orcid.org/0000-0001-7147-4535
U. V. Grygoriev https://orcid.org/0000-0002-2684-2586

References

  • [1] R.C. Tolman, Phys. Rev. C 55 (1939) 364.
  • [2] R.C. Tolman, Relativity, Thermodynamics, and Cosmology, (Dover Publications, New York, 1987; Oxford University Press, Oxford, 1934, 1946, 1949, 1987).
  • [3] L.D. Landau and E.M. Lifshitz, The Classical Theory of Fields, Course of Theoretical Physics, Vol. 2 (Butterworth-Heinemann, New York, 2003; FIZMATLIT, Moscow, 2003).
  • [4] J.R. Oppenheimer and G.M. Volkoff, Phys. Rev. C 55 (1939) 374.
  • [5] L.D. Landau and E.M. Lifshitz, Theoretical Physics, v.6 Fluid mechanics (Pergamon Press, 1987, FIZMATLIT, Moscow, 2013).
  • [6] S.L. Shapiro, S.A. Teukolsky, Black holes, white dwarfs, and neutron stars: The physics of compact objects (Wiley-VCH Verlag GmbH& Co.KGaA, Weinheim, 2004).
  • [7] G.S. Bisnovatiy-Kogan, Relativistic Astrophysics and Physical Cosmology (Moscow, Krasand, 2011) (in Russian); Stellar Physics: Stellar Evolution and Stability (Springer-Verlag, Berlin, Heidelberg, 2002).
  • [8] R.B. Wiringa, V. Fiks, A. Fabrocini, Phys. Rev. C 38 (1988) 1010.
  • [9] E. Chabanat, P. Bonche, P.Haensel, J. Meyer, and R. Shaeffer, Nucl. Phys. A 627 (1997) 710.
  • [10] A. Akmal, V.R. Pandharipande, and D.G. Ravenhall, Phys. Rev. C 58 (1998) 1804.
  • [11] I. Sagert, M. Hempel, C. Greiner, and J. Schaffner-Bielich, Eur, J. Phys. 27 (2006) 577.
  • [12] Bao-An Li, Lie-Wen Chen, and Che Ming Ko, Phys. Rep. 464 (2008) 113.
  • [13] N. Chamel and P. Haensel, Living Rev. Relativity 11 (2008) 10.
  • [14] A. Bauswein, S. Goriely, and H.-T. Janka, The Astrophysical Journal 773:78 1 (2013) (21pp), doi:10.1088/0004-637X/773/1/78.
  • [15] G. Giuliani, H. Zheng, A. Bonasera, Prog. Part. Nucl. Phys. 76 (2014) 116.
  • [16] R. Belverde, F. Cipolletta, C. Cherubini, S.M. de Carvalho, S. Filippi, R. Negreiros, J.P. Pereira, J.A. Rueda, and R. Ruffini. In: AIP Conference Proceedings; AIP Publishing, 2015; p. 030001-030001.19.
  • [17] Y. Lim and J.W. Holt, Eur.Phys. J. A 55 (2019) 209.
  • [18] Boyang Sun, Saketh Bhattiprolu, and James M. Lattimer, archiv:2311.00843v1 (2023).
  • [19] G. Baym, H. Bethe, and C.J. Pethick, Nucl. Phys. A 175 (1971) 225.
  • [20] P. Haensel, A.Y. Potehin, D.G. Yakovlev. Astrophysics and space science library, vol. 326, Neutron Stars 1. Equation of State and Structure (Springer, New York, 2007).
  • [21] J.M. Lattimer and M. Prakash, The astrophysical Journal 550 (2001) 426.
  • [22] C.J. Horowitz and J. Piekarewicz, Phys. Rev. Lett. 86 (2001) 5647.
  • [23] C.J. Horowitz and J. Piekarewicz, Phys. Rev. C 64 (2001) 062802(R).
  • [24] P.Haensel, Neutron Star Crusts N. Copernicus Astronomical Center, Polish Academy of Sciences, Bartycka 18, PL-00-716 Warszawa, Poland, 2000.
  • [25] Z. Arzoumanian, S. Bogdanov, J. Cordes, K. Gendreau, D. Lai, J. Lattimer, B. Link, A. Lommen, C. Miller, P. Ray, R. Rutledge, T. Strohmayer, C. Wilson-Hodge, and K. Wood, arXiv:0902.3264  , 2009; https://doi.org/10.48550/arXiv.0902.3264.
  • [26] A.F. Fantina, and F. Gulminelli, J. Phys., Conf. Ser. 2586 (2023) 012112; doi:10.1088/1742-6596/2586/1/012112 .
  • [27] H. Dinh, A.F. Fantina, and F. Gulminelli, Eur. Phys. J. A 59 (2023) 292; https://doi.org/10.1140/epja/s10050-023-01199-x .
  • [28] L. Wilets, Phys.Rev. 101 (1956) 1805; Rev. Mod. Phys. 30 (1958) 542.
  • [29] V.M. Strutinsky, and A.S. Tyapin, Exp. Theor. Phys. (USSR) 18 (1964) 664.
  • [30] A.S. Tyapin, Sov. Journ. Nucl. Phys. 11 (1970) 401; 13 (1971) 32; 14 (1972) 50.
  • [31] V.M. Strutinsky, A.G. Magner, and M. Brack, Z. Phys. A 319 (1984) 205.
  • [32] V.M. Strutinsky, A.G. Magner, and V. Yu. Denisov, Z. Phys. 322 (1985) 149.
  • [33] A.G. Magner, A.I. Sanzhur, and A.M. Gzhebinsky, Int. J. Mod. Phys. E 92 (2009) 064311.
  • [34] J.P. Blocki, A.G. Magner, P. Ring, and A.A. Vlasenko, Phys. Rev. C 87 (2013) 044304.
  • [35] J.P. Blocki, A.G. Magner, and P. Ring, Phys. Rev. C 92 (2015) 064311.
  • [36] J.S. Rowlinson and B. Widom, Molecular Theory of Capillarity (Clarendon Press, Oxford, 1982).
  • [37] H. Bethe, Annu. Rev. Nucl. Sci. 21 (1971) 93.
  • [38] M. Brack and R. K. Bhaduri, Semiclassical Physics (Addison-Wesley, Reading MA) 1997; 2nd edition (Westview Press, Boulder) 2003.
  • [39] M. Brack, G. Guet, H.-B. Håkansson, Phys. Rep. 123 (1985) 275.
  • [40] A. Bohr and B. Mottelson, Nuclear structure Vol. II (W. A. Benjamin, New York, 1975).
  • [41] A. G. Magner, V. M. Strutinsky, Z. Phys. A 324 (1985) 633.
  • [42] J.S. Rowlinson, Journ. Stat. Phys. 20 (1979) 197.
  • [43] D. Vautherin, D. Brink, Phys. Rev. C 5 (1972) 626.
  • [44] T.H.R. Skyrme, Philos. Mag. 1 8th ser. (1956) 1043.
  • [45] R.C. Barrett, D.F. Jacson, Nuclear sizes and structure (Oxford, Clarendon Press 1977).
  • [46] P. Ring, P. Schuck, The nuclear many-body problem (Berlin, Heidelberg, New York, Springer-Verlag 1980).
  • [47] J.P. Blaizot, Phys. Rep. 64 (1980) 172.
  • [48] H. Krivine, J. Treiner, and O. Bohigas, Nucl. Phys. A 336 (1980) 155.
  • [49] E. Chabanat, P. Bonche, P.Haensel, J. Meyer, and R. Shaeffer, Nucl. Phys. A 635 (1998) 231.
  • [50] B. Gramaticos, A. Voros, Ann. Phys. 123 (1979) 359; 129 (1980) 153.
  • [51] V.M. Kolomietz and A.I. Sanzhur, Eur. Phys. J. A 38 (2008) 345.
  • [52] V.M. Kolomietz, A.I. Sanzhur, and S. Shlomo, Phys. Rev. C 97 (2018) 064302.
  • [53] V.M. Kolomietz and S. Shlomo, Mean Field Theory (World Scientific, Singapore 2020).
  • [54] W.D. Myers and W.J. Swiatecki, Ann. Phys. (NY) 55 (1969) 395; 84 (1974) 186.
  • [55] W.D. Myers, W.J. Swiatecki, and C.S. Wang, Nucl.Phys. A 436 (1985) 185.
  • [56] P. Danielewicz and J. Lee, Int. J. Mod. Phys. E 18 (2009) 892.
  • [57] M. Centelles, X. Roca-Maza, X. Viñas, and M. Warda, Phys. Rev. Lett. 102 (2009) 122502.
  • [58] M. Warda, X. Viñas, X. Roca-Maza, and M. Centelles, Phys. Rev. C 80 (2009) 024316; 81 (2010) 054309; 82 (2010) 054314; arXiv:0906.0932 [nucl-th] (2009).
  • [59] X. Roca-Maza, M. Centelles, X. Viñas, and M. Warda, Phys. Rev. Lett. 106 (2011) 252501.
  • [60] X. Viñas, M. Centelles, X. Roca-Maza, and M. Warda, Eur. Phys. J. A 50 (2014) 27.
  • [61] J. Piekarewicz and M. Centelles, Phys. Rev.  C 79 (2009) 054311.
  • [62] T. Niksic, D. Vretenar, and P. Ring, Prog. Part. Nucl. Phys. 66 (2011) 519.
  • [63] B.D. Reed, F.J. Fattoyev, C.J. Horowitz, and J. Piekarewicz, Phys. Rev.  Lett. 126 (2021) 172503.
  • [64] N. N. Shchechilin, N. Chamel, J. M. Pearson, Phys. Rev. C 108 (2023) 025805.
  • [65] K.C. Chung and T. Kodama, Revesta Brasileira de Fisica 8 (1978) 404.
  • [66] V.M. Strutinsky, N.Ya. Ljashtchenko, N.A. Popov, Nucl.Phys. 46 (1963) 171.
  • [67] G. Raaijmaket et al., The Astrophysical Journ. Lett. 918:L29 (2021) 1 (13 pp).
  • [68] C.D. Capano et al., Nature Astronomy 4 (2020) 625.
  • [69] T.E. Riley et al., The Astrophysical Journ. Lett. 918:L27 (2021) 1 (30 pp).
  • [70] O. Lorenco, M. Bhuyan, C.H. Lenzi, M. Dutra, C. Gonzalez-Boquera, M. Centelles, X. Viñas, Phys. Lett. B 803 (2020) 135306.
  • [71] L.D. Landau and E.M. Lifshitz, Course of Theoretical Physics, v. 5 Statistical Physics (Third Edition Part 1, Elsevier, N.Y., 2005).
  • [72] B.K. Agrawal, S. Shlomo, and V.Kim Au, Phys. Rev. C 72 (2005) 014310.
  • [73] H. Zheng and A. Bonasera, Phys. Rev. C 83 (2011) 057602.