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

Systematic study of surface properties for Ne, Na, Mg, Al and Si isotopes in a coherent density fluctuation model using the relativistic mean field formalism

Jeet Amrit Pattnaik1 [email protected]    R. N. Panda1 [email protected]    M. Bhuyan2,3 [email protected]    S. K. Patra4,5 [email protected] 1Department of Physics, Siksha O{}^{\prime}O^{\prime} Anusandhan, Deemed to be University, Bhubaneswar-751030, India 2Center of theoretical and Computational Physics, Department of Physics, University of Malaya, Kuala Lumpur, 50603, Malaysia 3Institute of Research and Development, Duy Tan University, Da Nang 550000, Vietnam 4Institute of Physics, Sachivalya Marg, Bhubaneswar-751005, India 5Homi Bhabha National Institute, Training School Complex, Anushakti Nagar, Mumbai 400094, India
Abstract

We have systematically studied the surface properties, such as symmetric energy, neutron pressure, and symmetry energy curvature coefficient for Ne, Na, Mg, Al, and Si nuclei from the proton to neutron drip-lines. The Coherent Density Fluctuation Model (CDFM) is used to estimate these quantities taking the relativistic mean-field densities as inputs. The Brückner energy density functional is taken for the nuclear matter binding energy and local density approximation is applied for its conversion to coordinate space. The symmetry energy again decomposed to the volume and surface components within the liquid drop model formalism to the volume and surface parts separately. Before calculating the surface properties of finite nuclei, the calculated bulk properties are compared with the experimental data, whenever available. The NL3* parameter set with the BCS pairing approach in an axially deformed frame-work is used to take care of the pairing correlation when needed. The deformed density is converted to its spherical equivalent with a two Gaussian fitting, which is used as an input for the calculation of weight function in the CDFM approximation. With the help of the symmetric energy for the recently isotopes 29F, 28Ne, 29,30Na and 31,35,36Mg are considered to be within the island of inversion emphasized [Phys. Lett. B 772, 529 (2017)]. Although we get large symmetric energies corresponding to a few neutron numbers for this isotopic chain as expected, an irregular trend appears for all these considered nuclei. The possible reason behind this abnormal behavior of symmetry energy for these lighter mass nuclei is also included in the discussion, which gives a direction for future analysis.

pacs:
21.10.−k,21.10.Dr,21.10.Ft,21.10.Gv,21.60.−n,21.30.−x

I Introduction

The surface property of a nucleus is one of the pivotal quantities to determine the structural information of finite and infinite nuclear matter at isospin asymmetry. The symmetric SS energy determines the nuclear matter equation of state (EoS), which is connected with the isospin asymmetry. This SS controls the properties of the neutron star (NS), such as its mass and radius li2008 . Consequently, the Gravitational Waves (GW), the parameter of NS merger, etc. are automatically governed by the surface properties. Because of the interrelation of the nuclear matter symmetric energy with the neutron-skin thickness of finite nuclei, one can study the NS structure knowing the symmetric energy of the nuclear matter EoS li2019 . Similarly, there is a strong bond between the nuclear matter surface properties with the corresponding quantities of finite nuclei. There has been many significant efforts both in nuclear physics and astrophysics over the last two decades, and much progress in this regard is constraining both the magnitude of the symmetry energy S(ρ0)S(\rho_{0}) and the slope parameter LsymAL_{sym}^{A} = 3ρ0(dS(ρ)/dρ)ρ03\rho_{0}(dS(\rho)/d\rho)_{\rho_{0}} at the saturation density ρ0\rho_{0} of nuclear matter baran2005 ; li2008 ; tsang2012 ; horowitz2014 ; baldo2016 ; xu2020 .

Recently, a large number of works have been reported, which connect the surface properties of finite nuclei with the nuclear matter observables anto18 ; anto2 ; anto3 ; anto4 ; gad11 ; gad12 ; bhu18 ; bhu20 . To calculate these properties, a large number of methods are adopted such as Coherent Density Fluctuation Model anto1 ; anto2 ; anto3 ; anto4 ; gad12 ; bhu18 and Liquid Drop Approximation dani03 . Recently, the Coherent Density Fluctuation Model (CDFM) is gaining momentum for the evaluation of such properties. Although the original idea is published antozphys1980 long back, substantial work is waiting yet. Some of the relations of symmetric energy and related quantities have a strong correlation with the structures of the finite nuclei. We have reported in an isotopic series that the symmetric energy shows a peak at the corresponding magic number bhu18 ; bhu20 ; manpreet . Due to the advancement of the radioactive ion beam facility, the study of highly asymmetric isospin neutron/proton-rich nuclei open up a new area of research in Nuclear Physics. Some of the neutron number for certain proton combination behaves like a magic number. For example, neutron number N= 14, 16 and 32/34 act like close shell combination for neutron-rich Ne and Mg isotopes tanihata . The nuclei Ne, Na, Mg, Al, and Si show unexpected characteristics near the neutron-rich region. Some of the isotopic region termed as island of inversion. In this region of the mass table, the experimental binding energy deviates considerably from the shell model calculations. Long back Patra and Praharaj patra91 shown that these isotopes exhibit a large deformation in their ground states, which is later on verified by Tanihata et al. tanihata in their experimental study. In this paper, we would like to perform a systematic study of surface properties of these nuclei starting from the proton deficient region to the neutron-drip line and look for various structures.

A large number of theoretical formalisms are available in the literature for the structural analysis of finite nuclei and EoS of infinite nuclear matter. Among those well-known approaches, the Skyrme Hartree-Fock and relativistic mean-field (RMF) theory are the most successful frame-work for such study. The RMF approach has an advantage over the non-relativistic counterpart because of the inclusion of the spin-orbit interaction through the relativistic equations. This approach also reproduces the binding energy, quadruple deformation parameter β2\beta_{2}, and other related properties of finite nuclei throughout the mass table. With a few parameters of meson-nucleon couplings and their respective masses, the method successfully reproduces the experimental quantities for almost all known nuclei. This method also explains the nuclear EoS and NS properties quite well. Because of the satisfactory reproduction of experimental quantities by using RMF formalism, in the present paper, we use the formalism to get the bulk properties of finite nuclei including densities. These densities are used in the CDFM to generate surface properties of finite nuclei.

The paper is arranged as follows: The relativistic mean-field model and the fitting procedure of energy density functional (ρ){\cal{E}(\rho)} to an analytical expression in coordinate space with the help of the Brückner prescription bruk68 ; bruk69 is discussed in Sec. II. Also, the Coherent Density Fluctuation Model is briefly outlined. The Sec. III is dedicated to the discussion of our results derived from the calculations. A summary and concluding remarks are presented in Sec. IV.

II Relativistic mean field (RMF) Model

In this section, we briefly describe the formalism of the well documented relativistic mean-field theory. The standard nonlinear RMF Lagrangian density is built up by the interactions of isoscalar-scalar σ\sigma, isoscalar-vector ω\omega, and isovector-vector ρ\rho mesons with nucleons. In this version of the Lagrangian, the self-interaction of the σ\sigma-meson is also considered. Recently, the extended versions of the relativistic mean-field formalism based on the effective field theory motivated relativistic mean-field formalism are very popular frun96 ; frun97 . In these models, all possible meson-nucleon and their self-interactions are considered. The standard RMF formalism is quite successful for finite nuclei not only for β\beta-stable but also predicts reasonably the properties of drip-lines and super-heavy nuclei kumar18 ; kumar17 . Thus, we have considered this model in the present work. The RMF Lagrangian is discussed in detail in Refs. kumar18 ; kumar17 ; frun96 ; frun97 . During the last few decades, the pertinence of this decorum to nuclear astrophysics is in limelight. It foresees the structure of NS and reveals the tidal deformability adequately malik2018 . The recently reported experimental data of Gravitational Wave GW170817 measurements constraint the RMF nuclear EoS in a precise manner. Thus, the RMF Lagrangian gives enough confidence to adopt for further implementations to different nuclear environments. For a nucleon-meson interacting system the energy density functional is granted as kumar18 :

(r)\displaystyle{\cal E}({r}) =\displaystyle= α=p,nφα(r){i𝜶+β[MΦ(r)]+W(r)+12τ3R(r)+1+τ32A(r))}φα(r)\displaystyle\sum_{\alpha=p,n}\varphi_{\alpha}^{\dagger}({r})\Bigg{\{}-i\mbox{\boldmath$\alpha$}\!\cdot\!\mbox{\boldmath$\nabla$}+\beta\bigg{[}M-\Phi(r)\bigg{]}+W({r})+\frac{1}{2}\tau_{3}R({r})+\frac{1+\tau_{3}}{2}A({r})\bigg{)}\Bigg{\}}\varphi_{\alpha}(r) (1)
+(12+κ33!Φ(r)M+κ44!Φ2(r)M2)ms2gs2Φ2(r)12mω2gω2W2(r)12mρ2gρ2R2(r),\displaystyle\ +\left(\frac{1}{2}+\frac{\kappa_{3}}{3!}\frac{\Phi({r})}{M}+\frac{\kappa_{4}}{4!}\frac{\Phi^{2}({r})}{M^{2}}\right)\frac{m_{s}^{2}}{g_{s}^{2}}\Phi^{2}({r})-\frac{1}{2}\frac{m_{\omega}^{2}}{g_{\omega}^{2}}W^{2}({r})-\frac{1}{2}\frac{m_{\rho}^{2}}{g_{\rho}^{2}}R^{2}({r})\;,

Here, Φ\Phi, WW and RR are the re-considered fields for σ\sigma, ω\omega and ρ\rho mesons written as Φ=gsσ0\Phi=g_{s}\sigma_{0}, W=gωω0W=g_{\omega}\omega_{0} and R=gρρR=g_{\rho}\rho respectively. MM, mσm_{\sigma}, mωm_{\omega} and mρm_{\rho} are the nucleon masses ω\omega , σ\sigma and ρ\rho mesons, sequentially. From Eq. (1), we obtain our energy density nucl.{\cal{E}}_{nucl.} kumar18 ; kumar17 by taking that the exchange of mesons establish an uniform field, where the oscillations done by nucleons in a periodic motion said to be simple harmonic . From the effective-RMF energy density, the equation of motions (EoS) for the mesons and the nucleons are procured by using the Euler-Lagrange equation. A bunch of coupled differential equations are retained and settled accordingly kumar18 . The scalar and vector densities,

ρs(r)\displaystyle\rho_{s}(r) =\displaystyle= αφα(r)βφα,\displaystyle\sum_{\alpha}\varphi_{\alpha}^{\dagger}({r})\beta\varphi_{\alpha}, (2)
ρv(r)\displaystyle\rho_{v}(r) =\displaystyle= αφα(r)τ3φα,\displaystyle\sum_{\alpha}\varphi_{\alpha}^{\dagger}({r})\tau_{3}\varphi_{\alpha}, (3)

are figured out from the converged elucidations within the spherical harmonics.

II.1 Spherical Equivalent Density using two Gaussian Fitting

We have obtained the axially deformed density from RMF model for considered parameter sets and converted it into it’s spherical equivalent density by following the below steps,

  • Initially we have the deformed density ρ(r,z)\rho(r_{\bot},z) which is converted to one dimensional as ρ¯(ω)\bar{\rho}(\omega) where ω=x2+y2\omega=\sqrt{x^{2}+y^{2}} by performing the z-integration over the whole space as done in patra2009 ; panda2014 ; sharma2016 :

    ρ(ω)=ρ(ω2+z2)𝑑z.\vec{\rho}(\omega)=\int_{-\infty}^{\infty}\rho(\sqrt{\omega^{2}+z^{2}})dz\;. (4)
  • In the second step, the z-integrated density ρ¯(ω)\bar{\rho}(\omega) is fitted to a two Gaussian function expressed as:

    ρ(r)=i=12ciexp[air2],\rho(r)=\sum_{i=1}^{2}c_{i}exp[-a_{i}r^{2}]\;, (5)

    where the co-efficient cic_{i} and range aia_{i} are given with initial values for respective nuclei.

Then we have used the spherical equivalent density in place of deformed density for further calculations. The spherical equivalent densities are normalized to mass number of the nucleus.

To carry a detailed study of the nuclear bulk properties of open-shell nuclei, the contribution of pairing correlation becomes a major factor. In the present calculations, we will be using BCS approach for the calculation of pairing correlation. Here, the RMF+BCS formalism with NL3* parameter set lala09 is used for the finite nuclei. Within this approach, we evaluate the binding energy, quadruple deformation parameter β2\beta_{2}, root mean square radii and the axially deformed densities for both neutrons and protons distribution. The vector density ρv(r)\rho_{v}(r) is farther handled within the CDFM to locate the weight function |F(x)|2|F(x)|^{2}, which is an extensive quantity to reckon the symmetry energy (SAS^{A}), neutron pressure (PAP^{A}) and symmetry energy curvature coefficient (KsymAK_{sym}^{A}) for Ne, Na, Mg, Al and Si nuclei in the isotopic chains.

II.2 Nuclear Matter Parameter

The expression for the energy density of infinite and isotropic nuclear matter (NM) are obtained from the Brückner functional defined as bruk68 ; bruk69 :

(ρ)nucl.\displaystyle{\cal{E}}(\rho)_{nucl.} =\displaystyle= AV0(x)+VCVCx,\displaystyle AV_{0}(x)+V_{C}-V_{Cx}, (6)

where

V0(x)\displaystyle V_{0}(x) =\displaystyle= 37.53[(1+δ)5/3+(1δ)5/3]ρ0(x)2/3\displaystyle 37.53\Big{[}(1+\delta)^{5/3}+(1-\delta)^{5/3}]\rho_{0}(x)^{2/3} (7)
+b1ρ0(x)+b2ρ0(x)4/3+b3ρ0(x)5/3\displaystyle+b_{1}\rho_{0}(x)+b_{2}\rho_{0}(x)^{4/3}+b_{3}\rho_{0}(x)^{5/3}
+δ2[b4ρ0(x)+b5ρ0(x)4/3+b6ρ0(x)5/3].\displaystyle+\delta^{2}[b_{4}\rho_{0}(x)+b_{5}\rho_{0}(x)^{4/3}+b_{6}\rho_{0}(x)^{5/3}\Big{]}.

Here, b1=741.28b_{1}=-741.28, b2=1179.89b_{2}=1179.89, b3=467.54b_{3}=-467.54, b4=148.26b_{4}=148.26, b5=372.84b_{5}=372.84, b6=769.57b_{6}=-769.57. In each Flucton there are protons having Coulomb energy VC=35Z2e2xV_{C}=\frac{3}{5}\frac{Z^{2}e^{2}}{x} and Coulomb exchange energy VCx=0.7386Ze2(3Z/4πx3)1/3V_{Cx}=0.7386Ze^{2}(3Z/4\pi x^{3})^{1/3}.

The important part of the present calculation is to convert the NM quantities Eq. (6) from momentum (ρ\rho-) space to coordinate (rr-) space in local density approximation (LDA). The NM parameters SNMS^{NM}, PNMP^{NM}, LsymNML_{sym}^{NM} and KsymNMK_{sym}^{NM} are obtained from the well defined relations kumar18 ; Chen2014 ; bruk68 :

SNM\displaystyle S^{NM} =\displaystyle= 122(/ρ)α2|α=0,\displaystyle\frac{1}{2}\frac{\partial^{2}({\cal E}/\rho)}{\partial\alpha^{2}}\Big{|}_{\alpha=0}, (8)
LsymNM\displaystyle L_{sym}^{NM} =\displaystyle= 3ρS(ρ)ρ|ρ=ρ0=3PNMρ0,\displaystyle 3\rho\frac{\partial S(\rho)}{\partial\rho}\Big{|}_{\rho=\rho_{0}}=\frac{3P^{NM}}{\rho_{0}}, (9)
KsymNM\displaystyle K_{sym}^{NM} =\displaystyle= 9ρ22S(ρ)ρ2|ρ=ρ0.\displaystyle 9\rho^{2}\frac{\partial^{2}S(\rho)}{\partial\rho^{2}}\Big{|}_{\rho=\rho_{0}}. (10)

Here,

SNM\displaystyle S^{NM} =\displaystyle= 41.7ρ0(x)2/3+b4ρ0(x)+b5ρ0(x)4/3\displaystyle 41.7\rho_{0}(x)^{2/3}+b_{4}\rho_{0}(x)+b_{5}\rho_{0}(x)^{4/3} (11)
+b6ρ0(x)5/3,\displaystyle+b_{6}\rho_{0}(x)^{5/3},
PNM\displaystyle P^{NM} =\displaystyle= 27.8ρ0(x)5/3+b4ρ0(x)2+43b5ρ0(x)7/3\displaystyle 27.8\rho_{0}(x)^{5/3}+b_{4}\rho_{0}(x)^{2}+\frac{4}{3}b_{5}\rho_{0}(x)^{7/3} (12)
+53b6ρ0(x)8/3,\displaystyle+\frac{5}{3}b_{6}\rho_{0}(x)^{8/3},
KsymNM\displaystyle K_{sym}^{NM} =\displaystyle= 83.4ρ0(x)2/3+4b5ρ0(x)4/3\displaystyle-83.4\rho_{0}(x)^{2/3}+4b_{5}\rho_{0}(x)^{4/3} (13)
+10b6ρ0(x)5/3,\displaystyle+10b_{6}\rho_{0}(x)^{5/3},

are the nuclear matter quantities at local density. The nuclear densities of Ne, Na, Mg, Al and Si nuclei are calculated using RMF formalism. These densities are used as input in CDFM (described in the following sub-section) to calculate our weight function, which is a key quantity acting as a overpass among the NM parameters within (ρ\rho-) space and in (rr-) space of finite nuclei (manipulating LDA). To contest with the coordinate (rr-) and momentum (ρ\rho-) space simultaneously, we can build up the total density of a nucleus. The number of Fluctons in an infinite manner can be superimposed to form the total density of the nucleus, following the approach of CDFM discussed below.

II.3 Coherent Density Fluctuation Model

The Coherent Density Fluctuations Model (CDFM) is firstly prescribed by Antonov et al antozphys1980 , which plays a vital role to take care of the fluctuation of momentum and coordinate. This method can be easily used to interpret the surface properties of finite nuclei. In this CDFM formalism we can use NM tools SNMS^{NM}, PNMP^{NM} and KsymNMK_{sym}^{NM} from Eqs. (8)-(13) to obtain their values for a finite nucleus anto1 ; anto2 ; anto3 ; antozphys1980 . Within this model, the density ρ\rho (r, r) of a finite nucleus can be rewritten as the coherent superposition of infinite number of one-body density matrix (OBDM) ρx\rho_{x} (r, r) for spherical parts of NM coined as Fluctons bhu18 ; gad11 ,

ρx(𝐫)=ρ0(x)Θ(x|𝐫|),\rho_{x}({\bf r})=\rho_{0}(x)\,\Theta(x-|{\bf r}|), (14)

where ρo(x)=3A4πx3\rho_{o}(x)=\frac{3A}{4\pi x^{3}}. The generator coordinate x is the radius of a sphere consisting of Fermi gas having all the A nucleons distributed uniformly over it. It is appropriate to administer for such a system the OBDM disclosed as below bhu18 ; anto2 ; gad11 ; gad12 ,

ρ(𝐫,𝐫)=0𝑑x|F(x)|2ρx(𝐫,𝐫),\rho({\bf r},{\bf r^{\prime}})=\int_{0}^{\infty}dx|F(x)|^{2}\rho_{x}({\bf r},{\bf r^{\prime}}), (15)

where |F(x)|2|F(x)|^{2} is called as weight function (WF). The coherent superposition of OBDM ρx(𝐫,𝐫)\rho_{x}({\bf r},{\bf r^{\prime}}) is given below as:

ρx(𝐫,𝐫)\displaystyle\rho_{x}({\bf r},{\bf r^{\prime}}) =\displaystyle= 3ρ0(x)J1(kf(x)|𝐫𝐫|)(kf(x)|𝐫𝐫|)\displaystyle 3\rho_{0}(x)\frac{J_{1}\left(k_{f}(x)|{\bf r}-{\bf r^{\prime}}|\right)}{\left(k_{f}(x)|{\bf r}-{\bf r^{\prime}}|\right)} (16)
×Θ(x|𝐫+𝐫|2),\displaystyle\times\Theta\left(x-\frac{|{\bf r}+{\bf r^{\prime}}|}{2}\right),

where J1 is said to be the spherical Bessel function kind of first order and kfk_{f} is the Fermi momentum of nucleons inside the Flucton having radius xx and kf(x)=(3π2/2ρ0(x))1/3=γ/xk_{f}(x)=(3\pi^{2}/2\rho_{0}(x))^{1/3}=\gamma/x, (γ1.52A1/3\gamma\approx 1.52A^{1/3}). The Wigner distribution function of the OBDM of Eq. (16) is given by,

W(𝐫,𝐤)=0𝑑x|F(x)|2Wx(𝐫,𝐤).\displaystyle W({\bf r},{\bf k})=\int_{0}^{\infty}dx\,|F(x)|^{2}\,W_{x}({\bf r},{\bf k}). (17)

Here, Wx(𝐫,𝐤)=48π3Θ(x|𝐫|)Θ(kF(x)|𝐤|)W_{x}({\bf r},{\bf k})=\frac{4}{8\pi^{3}}\Theta(x-|{\bf r}|)\Theta(k_{F}(x)-|{\bf k}|). The density ρ\rho (r) in terms of weight function within the CDFM access is:

ρ(r)\displaystyle\rho(r) =\displaystyle= 𝑑𝐤W(𝐫,𝐤)\displaystyle\int d{\bf k}W({\bf r},{\bf k}) (18)
=0𝑑x|F(x)|23A4πx3Θ(x|𝐫|),\displaystyle=\int_{0}^{\infty}dx\,|F(x)|^{2}\,\frac{3A}{4\pi x^{3}}\Theta(x-|{\bf r}|),

which is normalized to A, i.e., ρ(𝐫)𝑑𝐫=A\int\rho({\bf r})d{\bf r}=A. In the δ\delta-function limit, the Hill-Wheeler integral equation, that is the differential equation for the WF in the generator coordinate is retrieved anto1 . The |F(x)|2|F(x)|^{2} for a provided density ρ\rho (r) is described as here

|F(x)|2=(1ρ0(x)dρ(r)dr)r=x,|F(x)|^{2}=-\left(\frac{1}{\rho_{0}(x)}\frac{d\rho(r)}{dr}\right)_{r=x}, (19)

with 0𝑑x|F(x)|2=1\int_{0}^{\infty}dx|F(x)|^{2}=1. A detailed genealogy can be found in Refs. bhu18 ; anto1 ; anto2 ; gad11 ; gad12 . The finite nuclear symmetry energy SAS^{A} neutron pressure PAP^{A} and surface curvature coefficients KsymAK_{sym}^{A} are calculated by weighting the corresponding quantities for infinite NM within the CDFM, as given below anto4 ; gad11 ; gad12 ; fuch95 ; anto17

SA=0𝑑x|F(x)|2SNM(ρ(x)),\displaystyle S^{A}=\int_{0}^{\infty}dx\,|F(x)|^{2}\,S^{NM}(\rho(x)), (20)
PA=0𝑑x|F(x)|2PNM(ρ(x)),\displaystyle P^{A}=\int_{0}^{\infty}dx\,|F(x)|^{2}\,P^{NM}(\rho(x)), (21)
KsymA=0𝑑x|F(x)|2KsymNM(ρ(x)).\displaystyle K_{sym}^{A}=\int_{0}^{\infty}dx\,|F(x)|^{2}\ K_{sym}^{NM}(\rho(x)). (22)

The SAS^{A}, PAP^{A} and KsymAK_{sym}^{A} in Eqs. (20\ref{s0}-22) are the surface weighted average of the corresponding NM quantities in the LDA limit for finite nuclei.

II.4 Volume and surface symmetric energy in Danielewicz’s liquid drop prescription

The nuclear binding energy E(A, Z) in the liquid drop model incorporating the volume symmetry energy parameter SVS_{V} and modified surface symmetry energy parameter SSS_{S} is written as steiner2005 ; myers :

E(A,Z)=B.A+ESA2/3+SVA(12Z/A)21+SSA1/3/SV\displaystyle E(A,Z)=-B.A+E_{S}A^{2/3}+S_{V}A\frac{(1-2Z/A)^{2}}{1+S_{S}A^{-1/3}/S_{V}}
+ECZ2A1/3+EdifZ2A+EexZ4/3A1/3+aΔA1/2.\displaystyle+E_{C}\frac{Z^{2}}{A^{1/3}}+E_{dif}\frac{Z^{2}}{A}+E_{ex}\frac{Z^{4/3}}{A^{1/3}}+a\Delta A^{-1/2}. (23)

In Eq. (23), B= binding energy per particle of symmetric NM at saturation. ESE_{S}, ECE_{C}, EdifE_{dif}, and EexE_{ex} are the coefficients of the surface energy of symmetric matter, the Coulomb energy, the diffuseness correction, and the exchange correction to the Coulomb energy, conjointly. The pairing corrections are delivered by the uttermost term, which is crucial for open-shell nuclei. The symmetry energy is rewritten as [3rd term of Eq. (23) in the form S = (NZ)2(N-Z)^{2}/A, where

S=SV1+SSSVA1/3=SV1+A1/3/κ.\displaystyle S=\frac{S_{V}}{1+\frac{S_{S}}{S_{V}}A^{-1/3}}=\frac{S_{V}}{1+A^{-1/3}/\kappa}. (24)

From the above Eq. (24), the individual components of SVS_{V} and SSS_{S} can be written as:

SV=S(1+1κA1/3)\displaystyle S_{V}=S\left(1+\frac{1}{\kappa A^{1/3}}\right) (25)

and

SS=Sκ(1+1κA1/3).\displaystyle S_{S}=\frac{S}{\kappa}\left(1+\frac{1}{\kappa A^{1/3}}\right). (26)

The symmetry energy and its volume and surface components are calculated within the CDFM formalism bhu18 ; anto2 ; antozphys1980 . Following Refs. dani03 ; dani04 ; dani06 , an approximate expression for the ratio κSVSS\kappa\equiv\frac{S^{V}}{S^{S}} can be written within the CDFM,

κ=3Rρ00𝑑x|F(x)|2xρ0(x)[(S(ρ0)S(ρ(x))1].\displaystyle\kappa=\frac{3}{R\rho_{0}}\int_{0}^{\infty}dx|F(x)|^{2}x\rho_{0}(x)\left[\left(\frac{S(\rho_{0})}{S(\rho(x)}\right)-1\right]. (27)

Here S(ρ0\rho_{0}) is the nuclear symmetry energy at equilibrium nuclear matter density ρ0\rho_{0}. Employing the density dependence of symmetry energy dani03 :

S[ρ(x)]=SV(ρ(x)ρ0)γ,\displaystyle S[\rho(x)]=S^{V}\left(\frac{\rho(x)}{\rho_{0}}\right)^{\gamma}, (28)

There exist various estimations for the value of the parameter γ\gamma. In present work, we use γ\gamma = 0.3 in reference to anto18 . Using above eq. and S(ρ0\rho_{0}) = SVS_{V}, the Eqs. (20) and (27) can be re-written as follows:

S=S(ρ0)0𝑑x|F(x)|2(ρ(x)ρ0)γ,\displaystyle S=S(\rho_{0})\int_{0}^{\infty}dx|F(x)|^{2}\left(\frac{\rho(x)}{\rho_{0}}\right)^{\gamma}, (29)

and

κ=3Rρ00𝑑x|F(x)|2xρ0(x)((ρ0ρ(x))γ1).\displaystyle\kappa=\frac{3}{R\rho_{0}}\int_{0}^{\infty}dx|F(x)|^{2}x\rho_{0}(x)\left(\left(\frac{\rho_{0}}{\rho(x)}\right)^{\gamma}-1\right). (30)
Refer to caption
Figure 1: (Color online) (a) The total binding energy B.E. (b) Binding energy per particle B.E./A and (c) Two neutron separation energy S2nS_{2n} as a function of neutron number N for Ne, Na, Mg, Al and Si isotopes, respectively.
Refer to caption
Figure 2: (color online) (a) The charge distribution radius and (b) the quadruple deformation parameter β2\beta_{2} as a function of neutron number N for Ne, Na, Mg, Al, and Si isotopes.

III Results and discussions

The surface properties, like symmetric energy, neutron pressure, and symmetry energy curvature coefficient of finite nuclei are studied using the coherent density fluctuation model by taking the RMF density as input in the calculations. Before going to the explanation of the surface properties of finite nuclei, we highlight herewith the bulk properties (binding energy, deformation, and nuclear charge radius) of the considered nuclei such as Ne, Na, Mg, Al, and Si. The results are discussed through Table 1 and Figs. 1-8. In Table 1 , the values of the coefficients of the Gaussian function [Eq. 5] c1c_{1}, c2c_{2} and a1a_{1}, a2a_{2} are given, which are used to get the spherical equivalent densities of the Ne, Na, Mg, Al and Si isotopes. The obtained spherical equivalent densities are further used to calculate the weight function |F(x)|2|F(x)|^{2} [Eq. 22]. Then the weight function |F(x)|2|F(x)|^{2} is used to calculate the nuclear matter parameters, such as symmetric energy SAS^{A}, neutron pressure PAP^{A} and symmetry energy curvature coefficient KsymAK_{sym}^{A} for finite nuclei. The predicted results are given in Figs. 1-8 in the following subsections.

III.1 Binding energy and two neutron-separation energy

To estimate the surface properties such as symmetry energy, neutron pressure, and symmetry energy curvature coefficient of finite nuclei, one needs to have an understanding of the ground state bulk properties. The binding energy (BE), root mean square (rms), charge radius (RchR_{ch}) and quadruple deformation parameter (β2\beta_{2}) for Ne, Na, Mg, Al, and Si isotopes starting from proton-rich to the expected neutron drip-line are obtained by using RMF model and NL3 parameter set. The calculated results with available experimental data for the above-considered isotopes are shown in Figures 1 and 2. From the overall observation of binding energy for all the considered isotopic chains, we note a good agreement between our results and the experimental data raman ; angeli . A careful analysis of Fig. 1 shows that the expected drip-lines for Ne, Na, Mg, Al, and Si isotopic chain are at mass number (A) = 34, 39, 40, 47, and 48 respectively.

Refer to caption
Figure 3: (color online) The neutron skin thickness rnp\triangle{r_{np}} is presented as a function of neutron number N for Ne, Na, Mg, Al and Si nuclei.
Refer to caption
Figure 4: (color online) The spherical equivalent density (solid line) for some of the selected nuclei are shown. The corresponding weight function (dashed line) are also displayed.

The two neutron separation energy S2nS_{2n} can be determined using the relation S2n(N,Z)S_{2n}(N,Z)= BE(N,Z) - BE(N-2,Z). Where both the BEsBE^{\prime}s are evaluated using RMF(NL3*) parameter for theory and experimental values are taken from the Ref. wang2017 for comparison. The estimated S2nS_{2n} values along with the available experimental data are presented in the lower panel of Fig. 1 for the 18-34Ne, 20-39Na, 22-40Mg, 24-47Al and 26-48Si isotopic series. One can notice from the figure, S2nS_{2n} decreases smoothly with an increase in neutron number except some sharp discontinuities (i.e kinks) at some particular neutron number NmagicN_{magic} indicating the shell closure property. It is worthy here to mention that the energy requires to remove two neutrons from a nucleus with NmagicN_{magic}+2 is much less than the nucleus with NmagicN_{magic} breaking the regular pattern of the S2nS_{2n}. In the lower panel of Fig. 2, we have displayed the total quadruple deformation parameter β2\beta_{2} for the considered isotopic series. It is clear from the figure that there are shape transitions at neutron numbers (N) = 14 for Ne, 20 for Na and Mg, 28 for Si isotopic series.

To find the ground state solution of a given nucleus, we obtain the conversed solutions by taking various initial quadruple deformation. From all those solutions, we have considered the maximum binding energy and corresponding quantities as the ground state of the given nucleus. The radial components, which are having odd under time-reversal symmetry and parity must be neglected. In the present paper, we have implemented the blocking method which rectifies the time-reversal symmetry. Normally we need to block different states around the Fermi level to determine the lowest energy states for the odd nucleus as per odd-odd, odd-even, and even-odd nuclei, the time-reversal symmetry violates. In order to take care of the time-reversal violation into account, we used the simple blocking procedure as discussed in Ref. patra . In many of the cases, we get almost nearly equal energies among various states indicating the shape co-existence nature of the isotopes. As we are following the criteria of taking the highest BE result, there are some low-laying excited states (degenerate states), which may be the reason behind the irregular plots. From those obtained values we got our ground state, excited states and degenerate states (almost same magnitude of binding energy with the different state).

In Fig. 2, we have displayed the charge distribution radius RchR_{ch} (upper panel) and quadruple deformation parameter β2\beta_{2} (lower panel) for the isotopic chain of Ne, Na, Mg, Al, and Si nuclei. The experimental data raman ; angeli are also compared, wherever available. Here in the lower mass region, the charge radius RchR_{ch} decreases and then increases with the mass number for all the isotopic chain. From the quadruple deformation parameter (β2\beta_{2}), we find shape transition appears at N = 14, 20, 20, 29, and 28 for Ne, Na, Mg, Al, and Si isotopes, respectively. It is to be noted that the irregular nature of the results mostly due to the shape co-existence nature of the isotopes. A low-lying excited state is very often noticed in these isotopic chains.

III.2 The neutron skin thickness

In Fig. 3, we have presented the neutron skin thickness rnp=rnrp\triangle r_{np}=r_{n}–r_{p} as a function of neutron number. The rnp\triangle r_{np} is an important quantity, which is connected with the surface property of the nucleus in terms of isospin asymmetry. It has a direct relation with the nuclear equation of state (EoS) which controls the structure of the neutron star. Thus, it has an important role in nuclear astrophysics. We have shown it for Ne, Na, Mg, Al, and Si isotopic series. As we see the neutron skin thickness increases with neutron number, i.e., the presence of more neutrons enhanced the nuclear radius of the nucleus. The neutron skin thickness in the nucleus is formed by a combination of volume and surface contributions. The volume part explains an increase in the local mean field of the surface of the neutron to the proton individual and the surface part in which the surface width of the neutron increases as compared to one of the individual protons. Both the contributions increase with neutron excess vinas2012 in an isotopic series. The linear increase in neutron skin thickness indicates the surface/volume saturation, which also tells the increase in symmetric energy.

III.3 The spherical equivalent density and weight function

In Fig. 4, we have plotted the spherical equivalent densities (solid) and corresponding weight functions |F(x)|2|F(x)|^{2} (dashed) as a function of radius. We have converted the deformed densities into their spherical equivalents through two Gaussian fittings by using Eq. 5. The spherical equivalent density for any nucleus can be easily obtained by using the values of c1c_{1}, a1a_{1} and c2c_{2}, a2a_{2} given in Table 1. For a careful inspection of Table 1, it is noticed that for all the nuclei, the values for c1c_{1} are found to be negative and c2c_{2} with a positive values. However, the values of a1a_{1} and a2a_{2} are almost comparable to each other with positive signs. More detail analysis of the fitting and also technical details can be found in Refs. patra2009 ; panda2014 ; sharma2016 .

Table 1: The axially deformed RMF density obtained from the NL3 parameter set are converted to its spherical equivalent using two Gaussian functionals. The coefficients c1c_{1}, a1a_{1} and c2c_{2}, a2a_{2} are listed for Ne, Na, Mg, Al, and Si isotopes.
Nucleus c1c_{1} a1a_{1} c2c_{2} a2a_{2} Nucleus c1c_{1} a1a_{1} c2c_{2} a2a_{2}
19Ne -3.16481 0.331586 3.47687 0.311655 20Ne -3.04522 0.333589 3.41182 0.313962
21Ne -3.59429 0.337037 3.95496 0.317189 22Ne -4.08107 0.339088 4.4364 0.319048
23Ne -3.83833 0.316714 4.17055 0.297795 24Ne -3.61795 0.42193 3.99949 0.373673
25Ne -1.61846 0.409952 2.0865 0.337241 26Ne -1.44537 0.400685 1.90112 0.322851
27Ne -0.856237 0.394508 1.30403 0.291229 28Ne -0.805704 0.365513 1.24565 0.274331
29Ne -2.09689 0.306084 2.50567 0.272813 30Ne -3.85888 0.276814 4.23455 0.260330
31Ne -2.18123 0.208782 2.44085 0.194627 32Ne -1.55223 0.202165 1.8495 0.187297
33Ne -1.07001 0.210559 1.37341 0.186965 34Ne -0.284251 0.253762 0.588779 0.171028
20Na -3.30838 0.33129 3.61359 0.309985 21Na -3.36721 0.336049 3.72673 0.315269
22Na -4.20169 0.339772 4.55721 0.320112 23Na -4.93147 0.342127 5.28423 0.323103
24Na -1.59925 0.3403 1.92878 0.289744 25Na -5.01558 0.302793 5.32092 0.287093
26Na -4.65003 0.297148 4.94264 0.280147 27Na -5.93258 0.398734 6.37662 0.369386
28Na -2.01555 0.388844 2.46011 0.324078 29Na -1.57596 0.361364 2.01826 0.297191
30Na -3.8188 0.315588 4.2291 0.291141 31Na -5.99753 0.27771 6.3129 0.263971
32Na -2.84109 0.216843 3.11021 0.203505 33Na -2.05982 0.217242 2.3596 0.200594
34Na -0.815459 0.235539 1.12203 0.193029 35Na -0.440892 0.262169 0.748431 0.182956
36Na -0.425172 0.257764 0.729539 0.178082 37Na -0.426312 0.252979 0.725345 0.174057
38Na -0.560764 0.236309 0.85135 0.174332 39Na -0.767664 0.221727 1.04873 0.174823
22Mg -3.5102 0.337634 3.86364 0.315119 23Mg -4.68847 0.341192 5.03992 0.32147
24Mg -5.58442 0.345115 5.93632 0.326273 25Mg -5.70374 0.32602 6.03196 0.308936
26Mg -5.91305 0.308559 6.21652 0.293159 27Mg -6.04836 0.305345 6.33983 0.289534
28Mg -6.28349 0.303037 6.56421 0.287016 29Mg -2.13344 0.407667 2.57616 0.332949
30Mg -3.19583 0.260324 3.48083 0.240888 31Mg -4.19839 0.228173 4.44081 0.216497
32Mg -2.65606 0.288664 2.97542 0.256755 33Mg -2.40436 0.311802 2.77146 0.270189
34Mg -1.27359 0.233869 1.56893 0.201356 35Mg -2.06917 0.231846 2.37601 0.209307
36Mg -0.808364 0.255092 1.1172 0.197461 37Mg -0.700842 0.25534 1.00767 0.191015
38Mg -0.779685 0.247165 1.07905 0.188717 39Mg -0.944354 0.236374 1.23856 0.187856
40Mg -1.6618 0.220155 1.94604 0.191281
24Al -1.45476 0.340458 1.78352 0.286481 25Al -5.33616 0.325916 5.6647 0.30783
26Al -8.96115 0.414522 9.43319 0.393303 27Al -4.25196 0.384866 4.61596 0.346254
28Al -9.60618 0.387342 10.0265 0.368222 29Al -10.1518 0.387245 10.5648 0.367966
30Al -8.57763 0.358694 8.99408 0.340275 31Al -3.93913 0.363267 4.39077 0.326436
32Al -2.62455 0.320816 3.00351 0.280775 33Al -1.90801 0.223972 2.13351 0.199603
34Al -1.76454 0.22558 2.02282 0.200253 35Al -4.40195 0.215828 4.66863 0.205308
36Al -4.45984 0.219293 4.74574 0.208377 37Al -4.54495 0.220995 4.83685 0.209645
38Al -4.7981 0.217558 5.07953 0.206567 39Al -4.69265 0.214125 4.97399 0.203054
40Al -3.60492 0.225556 3.88433 0.208409 41Al -3.50796 0.222836 3.78546 0.205275
42Al -2.99132 0.248402 3.32657 0.222474 43Al -3.05889 0.253024 3.39251 0.22516
44Al -3.02871 0.249435 3.36787 0.221941 45Al -2.9738 0.260693 3.31665 0.228223
46Al -3.1057 0.245182 3.43189 0.217474 47Al -0.319467 0.289022 0.63011 0.153346
26Si -7.91999 0.414491 8.39401 0.390892 27Si -4.17883 0.381022 4.53972 0.342842
28Si -3.26289 0.341632 3.63435 0.30699 29Si -3.778 0.363259 4.1903 0.326487
30Si -6.2779 0.265629 6.52777 0.253178 31Si -2.90607 0.336128 3.28085 0.294991
32Si -2.94072 0.303651 3.28119 0.270526 33Si -1.99993 0.20584 2.19687 0.186214
34Si -2.06805 0.212795 2.27827 0.191571 35Si -2.00258 0.217443 2.24468 0.195173
36Si -2.08563 0.206827 2.29686 0.185987 37Si -2.02117 0.221807 2.28665 0.197815
38Si -2.03642 0.226282 2.31387 0.200671 39Si -2.14661 0.225337 2.41648 0.199744
40Si -3.18035 0.30413 3.59945 0.266638 41Si -2.98979 0.250094 3.33998 0.225143
42Si -2.72873 0.236071 3.07376 0.21272 43Si -2.60821 0.230403 2.95853 0.20749
44Si -2.87398 0.234379 3.22239 0.211086 45Si -6.43192 0.225856 6.78428 0.215118
46Si -3.39586 0.244309 3.75371 0.219991 47Si -3.19051 0.231893 3.5316 0.208824
48Si -0.357803 0.322425 0.671824 0.159255
Refer to caption
Figure 5: (color online) The symmetric energy is shown as a function of neutron number N for Ne, Na, Mg, Al and Si nuclei.
Refer to caption
Figure 6: (color online) Here we have represented neutron pressure as a function of neutron number N for Ne, Na, Mg, Al and Si nuclei.

We have shown the equivalent density as a function of radial coordinate for 20,29Ne, 23,31Na, 24,39Mg and 28,45Si isotopes as representative cases. It is to be noted that the density plots in Fig. 4 reflect the normalized equivalent density of the deformed nucleus. In each panel, we have shown the equivalent densities for the isotopes from the β\beta-stable and neutron drip-line region for each atomic nucleus, which provides the information relative changes of the density with-respect-to isospin asymmetry. For example, panel (a), (b), (c), and (d) are assigned for the isotopes of 20,29Ne, 23,31Na, 24,30Mg, and 28,45Si, respectively. From the spherical equivalent density, we obtained the weight function |F(x)|2|F(x)|^{2} of the nucleus. This is one of the most important ingredients to determine the surface properties of the nucleus. In each specific representation, although the small difference in the surface region of the densities reflects a wide difference in the weight functions. As a consequence, one can expect a very different surface property for β\beta-stable and drip-line nuclei.

Refer to caption
Figure 7: (color online) The symmetry energy curvature coefficients are given for the Ne, Na, Mg, Al and Si nuclei isotopic series.
Refer to caption
Figure 8: (color online) The volume (SVS_{V}), surface (SsS_{s}) and total surface energy (S) are shown. The κ\kappa-factor is also given as a function of neutron number N for Ne, Na, Mg, Al and Si nuclei.

III.4 Nuclear surface properties

There are several prescriptions to analyze nuclear surface properties of finite nuclei in terms of isospin dependent nuclear matter quantities such as symmetry energy and its derivatives. In the present work, we have mainly highlighted the CDFM approach of Brückner’s functional and the liquid drop approximation of Danielewicz dani03 . Furthermore, the symmetry energy is separated into it’s volume and surface components, which are mainly approximated in connection with the nuclear matter’s properties at the saturation. For example, Eq. (29) shows that the saturation symmetric energy S(ρ0)S(\rho_{0}) and nuclear matter density are needed to calculate the symmetric energy of a finite nucleus. On the other hand, for the same quantity, we need in the nuclear matter Brückner’s functional as of Eqs. (20)-(22). The calculated nuclear surface properties such as symmetric energy, neutron pressures, and symmetry energy curvature are given in Fig. 5, 6, and 7, respectively as a function of the neutron number. In each figure, we have given the corresponding calculated quantities for the isotopic chain of considered nuclei Ne, Na, Mg, Al, and Si. From the Fig. 5, it can be noticed that the symmetry energy for all the isotopes bounded within a range of 23-25 MeVMeV. It is worth mentioning that the exotic nuclei having large neutron-to-proton asymmetry respond to the nature of nuclear symmetry energy. Further, the appropriate information of symmetry energy from finite nuclei will be added to a wide range of nuclear phenomena starting from the study of nuclear structure, dynamics of heavy-ion reactions to the high isospin asymmetry system such as neutron star matter.

Observing the characteristics of symmetry energy SAS^{A}, and surface symmetry curvature KsymAK_{sym}^{A} in Figs. 5, and 7, in general, distorted peaks appear at N = 14, 16, 20, and 28, for some of the cases. Similarly, we also notice a fall in the neutron pressure PAP^{A} at the same neutron numbers in Fig. 6. We found the value of SAS^{A} within the range of 24-25 MeVMeV at the neutron number N = 20 for all the isotopic series of Ne, Na, Mg, Al, and Si nuclei. In all the figures, the trend is not smooth, this is either from the transform of deformed density into the spherical equivalent one by using the two Gaussian fittings (discussed in sub-section III.3) or by neglecting the shape degrees of freedom. Although there is an anomalous trend appear in all the isotopic chain, a careful inspections show that the magnitude of the peak and/or depth are assuredly at the neutron magic N =20 and also for 28 along with a few more neutron numbers, namely N = 14, and 16. These larger (lower) magnitude of SAS^{A}, PAP^{A} and KsymAK_{sym}^{A} provides the signature of shell/sub-shell closures over the isotopic chain. It is to be noted that, one can obtain the KsymAK_{sym}^{A} by adopting the Thomas-Fermi approach along with relativistic mean-field theory within the leptodermus expansion for finite nuclei patra2002curv . Here the neutron pressure and symmetry energy curvature are distributed within a range of -7 to -1 MeV/fm3MeV/fm^{3} and -550 to -250 MeVMeV respectively. In some of our earlier study bhu18 ; abdul19 ; manpreet , we notice a relation of SAS^{A} with the shell closure nature of magic number, i.e., we get a maximum of SAS^{A} at the magic number. This correlation is seen only for light and medium mass nuclei. While searching such relation in the considered mass region although, we noticed such type of peaks at the presently reported magic number plb2020Han within the Brückner’s functional, but it does not show a regular pattern, as we are getting in medium mass region manpreet . For further analysis of the surface properties of symmetry energy, we subdivided the SS values into its surface and volume components as shown in Fig. 8. To calculate these quantities, we have used the prescription of P. Danielewicz dani03 ; dani04 ; dani06 ; dani09 . The estimation of SS with Eq. (29) is quite different than that of (20) dani03 ; dani04 ; dani06 ; dani09 . As a result, we predict nearly two different values of SS in both two approaches, which can be seen clearly in Figs. (5) and (8). In this Fig. 8 we have produced the volume, the surface, and the total surface energy. The parameter κ=SV/SS\kappa=S_{V}/S_{S}, i.e., the ratio of the volume to the surface part of the symmetric energy is also given in the figure. We observed that the symmetry energy related to volume is more than the surface contribution and their ratio κ\kappa decreases with increasing of N. As the symmetric energy is pretty much related to N-Z asymmetry, hence it is crucial to examine the surface and volume part. At the surface of the nucleus, the nuclear saturation does not acquire, and also due to the unequal population of the number density of the nucleons, the surface symmetry energy get more crucial, detail explanation can be found in Ref. feenberg1947 .

IV Concluding Remarks

In our outline, the properties of the surface of a nucleus, such as symmetric energy, neutron pressure, and the symmetry energy curvature coefficient as a function of the neutron number are investigated for Ne, Na, Mg, Al, and Si nuclei from the proton to neutron drip-lines. The Coherent Density Fluctuation Model (CDFM) is used to evaluate these quantities within the framework of the relativistic mean-field formalism. Here, in CDFM the nuclear/nucleonic density does not follow any sharp edge surface; thus, the diffuseness parameter is also not neglected. The fluctuation of the flucton attains infinitesimal size/dimension and is distributed over the range of nuclear density distribution from the relativistic mean-field model. Hence, there is no possibility of sloppy surface contribution from the density in terms of weight function. And further, we have examined the CDFM account the surface effect substantially compared to Liquid-Drop-Approximation subrat and reference therein. We have discussed the bulk properties of these isotopic chain, which are essential for determining the symmetry energy and its co-efficient. We evaluated the binding energy, the binding energy per particle, charge radius, nuclear quadruple moment deformation parameter, and the neutron-skin thickness. To locate the recently reported magic structure of some of the isotopes in the considered region, we analyzed the two-neutron separation energy. The correlation of the magic number with the surface symmetric properties of the isotopes are studied. We noticed some relation in the reported isotopes, which is recently a fact. As expected, the neutron-skin thickness increases monotonously with neutron numbers in an isotopic series. On the other hand, the BE/A increases initially and then decreases with N. The shape of the nucleus in an isotopic chain changes and does not show any regular pattern. This is because a larger number of nuclei exhibit shape coexistence.

We converted the deformed densities to their spherical equivalent with a two Gaussian fitting. The coefficients a1a_{1} and a2a_{2} are in general decrease, but c1c_{1} increases in an zig-zag manner. The values of c2c_{2} however decrease with neutron number. In practice, when we use the Gaussian fitting for the axial density distribution to obtain the spherical one, there is an enhancement of central density with a long tail. As a result, the surface covers a wide range of the radius, which is the limitation of the Gaussian fitting procedure. In some of the previous studies by our collaborators, various fitting techniques were used to convert axial deform density to spherical form and their effect on the surface properties of finite nuclei quantitatively kishore20 and reference therein.

In our present work, we reported here two types of symmetry energies, one due to liquid drop approximation and the other due to LDA by Brückner method. Although we can not measure the symmetric energy often directly, we can extract the values from other nuclear observable, indirectly which are related to it bhu18 . But the question is why we are not getting the same values or the same trends of symmetry energy for two different approximations? Although the dimensions are the same in both approaches the expressions are different and the fact is that these nuclei are small in size. Nuclei having large size show almost the same values of symmetry energy in both approximations as we have shown in our previous work quddus2020 . Here we get the maximum value of symmetry energy in the liquid drop approach which is 33.1133.11 and 25.5225.52 MeVMeV in Brückner functional adoption for NL3* parameter set.

In the earlier studies, it is shown that the symmetric energy and/or neutron pressure is connected with the shell/sub-shell closure in the emergence of a peak and/or fall over an isotopic chain gad11 ; gad12 ; bhu18 ; quddus2020 . Hence the symmetry energy and its co-efficient (neutron pressure) can be used as an observable for predicting the neutron and/or proton magic at the drip-line region. Although we get considerable magnitude in the symmetric energy for these neutron numbers (N=14, 16 and 32/34), but a broke-down/ distorted trend in the isotopic chain of all the considered nuclei. The abnormal trend in these mass regions can be correlated with the possible reasons:

  • The considered nuclei in the present study are lighter mass nuclei, hence quite sensitive to the neutron-to-proton ratio. Analogous to the neutron proton ratio, the neutron-to-proton difference is very small, hence the isospin dependency of the system not effective, as a result, the symmetry energy for these mass nuclei is not followed the systematic trend.

  • The considered mass nuclei are found to be highly deformed in their ground state, which reflects in their densities. Following the limitation of the calculation, the spherical equivalent densities using a slender approximation are implemented into the CDFM to obtain the weight function. It is worth mentioning that, the density in terms of weight function is one of the central ingredients for obtaining the symmetry energy kishore20 and its components.

  • As we mentioned above, the considered isotopes are small in size, hence it is likely to get clustering configuration with sharp surface instead of a smooth surface as that of nuclei of mass A \geq 40 arima72 ; aremugumpatra05 ; bhu13 ; nora08 , it has been shown that the smaller nuclei in the mass table show cluster properties unlike developing a smooth surface as that of medium or heavy mass nuclei.

Therefore we suggest, while doing these further calculations in the lower mass region, one must need to take care of these points.

Acknowledgement: One of the authors (JAP) is thankful to the Institute of Physics, Bhubaneswar for providing the computer facilities during the work. This work is partly reinforced by SERB, Department of Science and Technology, Govt. of India, Project No. CRG/2019/002691. MB acknowledges the support from FAPESP Project No. 2017/05660-0, FOSTECT Project No. FOSTECT.2019B.04, and the CNPq - Brasil.

References

  • (1) B. A. Li, L. W. Chen, and C. M. Ko, Phys. Rep., 464, 113 (2008).
  • (2) LI, Bao-An, krastev, Plamen G., Wen, and Zhang, Nai-Bo, Eur. Phys. J. A, 55, 117 (2019).
  • (3) V. Baran, M. Colonna, V. Greco, and M. Di Toro, Phys. Rep., 410, 335 (2005).
  • (4) M. B. Tsang et al., Phys. Rev. C, 86, 015803 (2012).
  • (5) C. J. Horowitz, E. F. Brown and Y. Kim and others, J. Phys. G, 41, 093001 (2014).
  • (6) M. Baldo and G. F. Burgio, Prog. Part. Nucl. Phys., 91, 203 (2016).
  • (7) Jun Xu, Wen-Jie Xie, and Baon- An Li, Phys. Rev. C, 102, 044316 (2020).
  • (8) A. N. Antonov, D. N. Kadrev, and M. K. Gaidarov et al., Phys. Rev. C, 98, 054315 (2018).
  • (9) A. N. Antonov, V. A. Nikolaev, and I. Z. Petkov, Z. Phys. A At. and Nucl. 297, 297 (1980)
  • (10) A. N. Antonov, V. A. Nikolaev, and I. Z. Petkov, Z. Phys. A At. and Nucl., 304, 239 (1982).
  • (11) A. N. Antonov, V. A. Nikolaev, and I. Z. Petkov, Nuovo Ci-mento A, 86, 23 (1985).
  • (12) A. N. Antonov, D. N. Kadrev, and P. E. Hodgson, Phys. Rev. C, 50, 164 (1994).
  • (13) M. K. Gaidarov, A. N. Antonov, and others, Phys. Rev. C, 84, 034316 (2011).
  • (14) M. K. Gaidarov, A. N. Antonov, and others, Phys. Rev. C, 85, 064319 (2012).
  • (15) M. Bhuyan, B. V. Carlson, S. K. Patra, and S.-G. Zhou, Phys. Rev. C, 97, 024322 (2018).
  • (16) M. Bhuyan, R. Kumar, S. Rana, D. Jain, S. K. Patra, and B. V.Carlson, Phys. Rev. C, 101, 044603 (2020).
  • (17) P. Danielewicz, Nucl. Phys. A, 727, 233 (2003).
  • (18) A. Antonov, V. Nikolaev, and I. Petkov, Zeitschrift für PhysikA At. and Nucl., 297, 257 (1980).
  • (19) M. Kaur, A. Quddus, A. Kumar, M. Bhuyan, and S. K. Patra, J. of Phys. G: Nucl. and Part. Phys., 47, 105102 (2020).
  • (20) I. Tanihata, D. Hirata, T. Kobayashi, S. Shimoura, K. Sugimoto,and H. Toki, Phys. Lett. B, 289, 261 (1992).
  • (21) S. K. Patra and C. Praharaj, Phys. Lett. B, 273, 13 (1991).
  • (22) K. A. Brueckner, J. R. Buchler, S. Jorna, and R. J. Lombard,Phys. Rev., 171, 1188 (1968).
  • (23) K. A. Brueckner, J. R. Buchler, R. C. Clark, and R. J. Lombard, Phys. Rev., 181, 1543 (1969).
  • (24) R. Furnstahl, B. D. Serot, and H.-B. Tang, Nucl. Phys. A, 598, 539 (1996).
  • (25) R. Furnstahl, B. D. Serot, and H.-B. Tang, Nucl. Phys. A, 615, 441 (1997).
  • (26) B. Kumar, S. K. Patra, and B. K. Agrawal, Phys. Rev. C, 97, 045806 (2018).
  • (27) B. Kumar, S. Singh, B. Agrawal, and S. K. Patra, Nucl. Phys. A, 966, 197 (2017).
  • (28) T. Malik, N. Alam, M. Fortin, et al., Phys. Rev. C, 98, 035804 (2018).
  • (29) B. P. Abbott, R. Abbott, T. D. Abbott,et al.(LIGO Scientific Collaboration and Virgo Collaboration), Phys. Rev. Lett., 119, 161101 (2017).
  • (30) S. K. Patra, R. N. Panda, P. Arumugam, and R. K. Gupta, Phys. Rev. C, 80, 064602 (2009).
  • (31) R. N. Panda, M. K. Sharma, and S. K. Patra, Mod. Phys. Lett. A, 29, 1450013 (2014).
  • (32) M. K. Sharma, R. N. Panda, M. K. Sharma, and S. K. Patra, Phys. Rev. C, 93, 014322 (2016).
  • (33) G. Lalazissis, S. Karatzikos, R. Fossion, D. P. Arteaga, A. Afanasjev, and P. Ring, Phys. Lett. B, 671, 36 (2009).
  • (34) W.-C. Chen and J. Piekarewicz, Phys. Rev. C, 90, 044305 (2014).
  • (35) C. Fuchs, H. Lenske, and H. H. Wolter, Phys. Rev. C, 52, 3043 (1995).
  • (36) A. N. Antonov, D. N. Kadrev, M. K. Gaidarov, P. Sarriguren, and E. M. de Guerra, Phys. Rev. C, 95, 024314 (2017).
  • (37) A. Steiner, M. Prakash, J. Lattimer, and P. Ellis, Phys. Rep., 411, 325 (2005).
  • (38) W. D. Myers and W. J. Swiatecki, Nucl. Phys., 81, 1 (1966).
  • (39) P. Danielewicz, (2004), arXiv:nucl-th/0411115.
  • (40) P. Danielewicz, “Symmetry energy,” in Opportunities with Exotic Beams(2007) pp. 142–151.
  • (41) S. Raman, C. Nestor, S. Kahane, and K. Bhatt, Atom. Data and Nucl. Data Tables, 42, 1 (1989).
  • (42) I. Angeli and K. P. Marinova, Atom. Data and Nucl. Data Tables, 99, 69 (2013).
  • (43) M. Wang, G. Audi, F. G. Kondev, W. J. Huang, N. S., and X. Xing, Chin. Phys. C, 41, 030003 (2017).
  • (44) S. K. Patra, M. Del Estal, M. Centelles, and X. Vïnas, Phys. Rev. C, 63, 024311 (2001).
  • (45) X. Vïnas, M. Warda, M. Centelles, and X. Roca-Maza, Int. J. of Mod. Phys. E, 21, 1250029 (2012).
  • (46) S. K. Patra, M. Centelles, X. Vïnas, and M. Del Estal, Phys. Rev. C, 65, 044304 (2002).
  • (47) A. Quddus, M. Bhuyan, and S. K. Patra, J. of Phys. G:Nucl. and Part. Phys., 47, 045105 (2020).
  • (48) R. Han, X. Li, W. Jiang, et al., Phys. Lett. B, 772, 529 (2017).
  • (49) P. Danielewicz and J. Lee, Int. J. of Mod. Phys. E, 18, 892 (2009).
  • (50) E. Feenberg, Rev. Mod. Phys., 19, 239 (1947).
  • (51) S. K. Biswal, M. K. Abu El Sheikh, N. Biswal, and others, Nucl. Phys. A 1004, 122042 (2020).
  • (52) A. Quddus, M. Bhuyan, and S. K. Patra, J. of Phys. G: Nucl. and Part. Phys., 47, 045105 (2020).
  • (53) K. C. Naik, M. Kaur, A. Kumar, and S. K. Patra, Int. J. of Mod. Phys. E, 28, 1950100 (2019).
  • (54) A. Arima, Horiuchi,et al., “Clustering in light nuclei,” in Adv. in Nucl. Phys.: Volume 5, edited by M. Barangerand E. Vogt (Springer US, Boston, MA, 1972) pp. 345–477.
  • (55) P. Arumugam, B. K. Sharma, S. K. Patra, and R. K. Gupta, Phys. Rev. C, 71, 064308 (2005).
  • (56) M. Bhuyan, Int. J. of Mod. Phys. E, 22, 1350068 (2013).
  • (57) N. Furutachi, M. Kimura,et al., Prog. of Theo. Phys., 119, 403 (2008).