arxiv: v1 [nucl-th] 19 Nov 2014

Similar documents
Pygmy dipole resonances in stable and unstable nuclei

The Nuclear Equation of State

Density dependence of the nuclear symmetry energy estimated from neutron skin thickness in finite nuclei

Theory of neutron-rich nuclei and nuclear radii Witold Nazarewicz (with Paul-Gerhard Reinhard) PREX Workshop, JLab, August 17-19, 2008

Density dependence of the nuclear symmetry energy estimated from neutron skin thickness in finite nuclei

Probing the Nuclear Symmetry Energy and Neutron Skin from Collective Excitations. N. Paar

Nuclear Matter Incompressibility and Giant Monopole Resonances

arxiv: v1 [nucl-th] 21 Aug 2007

Nature of low-energy dipole states in exotic nuclei

Low-lying dipole response in stable and unstable nuclei

Relativistic versus Non Relativistic Mean Field Models in Comparison

Nuclear Landscape not fully known

arxiv: v2 [nucl-th] 28 Aug 2014

Mean-field concept. (Ref: Isotope Science Facility at Michigan State University, MSUCL-1345, p. 41, Nov. 2006) 1/5/16 Volker Oberacker, Vanderbilt 1

Nuclear symmetry energy deduced from dipole excitations: comparison with other constraints

An empirical approach combining nuclear physics and dense nucleonic matter

Observables predicted by HF theory

Investigation of Nuclear Ground State Properties of Fuel Materials of 232 Th and 238 U Using Skyrme- Extended-Thomas-Fermi Approach Method

Temperature Dependence of the Symmetry Energy and Neutron Skins in Ni, Sn, and Pb Isotopic Chains

Temperature Dependence of the Symmetry Energy and Neutron Skins in Ni, Sn, and Pb Isotopic Chains

Isoscalar dipole mode in relativistic random phase approximation

Nuclear Equation of State from ground and collective excited state properties of nuclei

Quantum Theory of Many-Particle Systems, Phys. 540

What did you learn in the last lecture?

Collective excitations in nuclei away from the valley of stability

Evolution Of Shell Structure, Shapes & Collective Modes. Dario Vretenar

Structure properties of medium and heavy exotic nuclei

Preliminary Studies of Thermal Wavelength Approximation in 208 Pb and 91 Zr hot Nuclei

The Nuclear Many-Body Problem

Density dependence of the symmetry energy and the nuclear equation of state : A dynamical and statistical model perspective

Shape of Lambda Hypernuclei within the Relativistic Mean-Field Approach

Discerning the symmetry energy and neutron star properties from nuclear collective excitations

Dipole Response of Exotic Nuclei and Symmetry Energy Experiments at the LAND R 3 B Setup

Neutron star structure explored with a family of unified equations of state of neutron star matter

Constraining the symmetry energy based on relativistic point coupling interactions and excitations in finite nuclei

1 Introduction. 2 The hadronic many body problem

arxiv: v1 [nucl-th] 26 Jun 2011

Covariance analysis for Nuclear Energy Density Functionals

Fusion Barrier of Super-heavy Elements in a Generalized Liquid Drop Model

Stability of heavy elements against alpha and cluster radioactivity

The IC electrons are mono-energetic. Their kinetic energy is equal to the energy of the transition minus the binding energy of the electron.

Dipole Polarizability and the neutron skin thickness

arxiv:nucl-th/ v2 4 Apr 2003

in covariant density functional theory.

Coupling of giant resonances to soft E1 and E2 modes in 8 B

Symmetry energy, masses and T=0 np-pairing

Symmetry energy of dilute warm nuclear matter

Gianluca Colò. Density Functional Theory for isovector observables: from nuclear excitations to neutron stars. Università degli Studi and INFN, MIlano

The effect of the tensor force on the predicted stability of superheavy

Nuclear symmetry energy and Neutron star cooling

Dipole Polarizability and Parity Violating Asymmetry in 208 Pb

Pb, 120 Sn and 48 Ca from High-Resolution Proton Scattering MSU 2016

Density dependence of the nuclear symmetry energy estimated from neutron skin thickness in finite nuclei

Nuclear collective vibrations in hot nuclei and electron capture in stellar evolution

Tamara Nikšić University of Zagreb

Relativistic Hartree-Bogoliubov description of sizes and shapes of A = 20 isobars

SEMICLASSICAL APPROACHES TO THE COUPLINGS BETWEEN NUCLEAR DIPOLE MODES AND SURFACE VIBRATIONS

Localized form of Fock terms in nuclear covariant density functional theory

Volume and Surface Components of the Nuclear Symmetry Energy

Compressibility of Nuclear Matter from Shell Effects in Nuclei

Relativistic Radioactive Beams as a Tool for Nuclear Astrophysics

Neutron Halo in Deformed Nuclei

Equation-of-State of Nuclear Matter with Light Clusters

Clusters in Dense Matter and the Equation of State

RFSS: Lecture 2 Nuclear Properties

Beyond mean-field study on collective vibrations and beta-decay

Reactions of neutron-rich Sn isotopes investigated at relativistic energies at R 3 B

Equation of State of Dense Matter

arxiv: v1 [nucl-th] 14 Feb 2012

Statistical Behaviors of Quantum Spectra in Superheavy Nuclei

arxiv: v1 [nucl-th] 29 Jun 2009

Correction to Relativistic Mean Field binding energy and N p N n scheme

Shape Coexistence and Band Termination in Doubly Magic Nucleus 40 Ca

Thermodynamics of nuclei in thermal contact

Mean-field description of fusion barriers with Skyrme s interaction

The pygmy dipole strength, the neutron skin thickness and the symmetry energy

Relativistic mean-field description of collective motion in nuclei: the pion field

Isospin asymmetry in stable and exotic nuclei

The role of isospin symmetry in collective nuclear structure. Symposium in honour of David Warner

Physics 228 Today: April 22, 2012 Ch. 43 Nuclear Physics. Website: Sakai 01:750:228 or

Krane Enge Cohen Willaims NUCLEAR PROPERTIES 1 Binding energy and stability Semi-empirical mass formula Ch 4

Chapter 6. Isospin dependence of the Microscopic Optical potential for Neutron rich isotopes of Ni, Sn and Zr.

TWO CENTER SHELL MODEL WITH WOODS-SAXON POTENTIALS

Direct reactions methodologies for use at fragmentation beam energies

Correlating the density dependence of the symmetry y energy to neutron skins and neutron-star properties

The dipole strength: microscopic properties and correlations with the symmetry energy and the neutron skin thickness

Some new developments in relativistic point-coupling models

arxiv:nucl-th/ v1 27 Apr 2004

Nuclear Mean Fields through Selfconsistent Semiclassical Calculations

Fermi-Liquid Theory for Strong Interactions

Nuclear matter inspired Energy density functional for finite nuc

arxiv: v1 [nucl-th] 16 Jan 2019

An extended liquid drop approach

THE PYGMY DIPOLE CONTRIBUTION TO POLARIZABILITY: ISOSPIN AND MASS-DEPENDENCE

Microscopic Fusion Dynamics Based on TDHF

Lisheng Geng. Ground state properties of finite nuclei in the relativistic mean field model

Surface energy coefficient determination in global mass formula from fission barrier energy Serkan Akkoyun 1,* and Tuncay Bayram 2

Spin Cut-off Parameter of Nuclear Level Density and Effective Moment of Inertia

Investigation of the Giant Monopole Resonance in the Cd and Pb Isotopes: The Asymmetry Term in Nuclear Incompressibility and the MEM Effect

Rising the of pygmy dipole resonance in nuclei with neutron excess

Transcription:

Derivative corrections to the symmetry energy and the isovector dipole-resonance structure in nuclei arxiv:1411.5183v1 [nucl-th] 19 Nov 214 Abstract J P Blocki National Centre of Nuclear Research, PL-681 Warsaw, Poland A G Magner Institute for Nuclear Research, Kyiv 368, Ukraine P Ring Technical Munich University, D-85747 Garching, Germany The effective surface approximation is extended accounting for derivatives of the symmetry energy density per particle. Using the analytical isovector surface energy constants within the Fermi-liquid droplet model, one obtains energies and sum rules of the isovector dipole resonance structure in a reasonable agreement with the experimental data and other theoretical approaches. Pacs Ref: 21.1.Dr, 21.65.Cd, 21.6.Ev, 24.3.Cz Keywords: Nuclear binding energy, liquid droplet model, extended Thomas-Fermi approach, surface symmetry energy, neutron skin, isovector stiffness. 1. Introduction The symmetry energy is a key quantity for study of the fundamental properties of exotic nuclei with a large excess of neutrons above protons in the nuclear and astronomic physics [1, 2, 3, 4, 5, 6, 7, 8]. In spite of a very intensive study of these properties, the derivatives of the symmetry energy and its surface coefficient are still rather undetermined in the calculations by the liquid droplet model (LDM) [1], or within more general local density approach (LDA) [9, 1], in particular within the extended Thomas- Fermi (ETF) approximation[11], and models based on the Hartree-Fock (HF) method [12] with applying the kyrme forces [13, 14, 15, 16], in contrast to the basic volume symmetry energy constant. Within the nuclear LDA, the variational condition derived from minimizing the nuclear energy at the fixed particle number and the neutron excess above protons can be simplified using the expansion in a small parameter a/r A 1/3 for heavy enough nuclei with a being of the order of the diffuse edge thickness of the nucleus, R the mean curvature radius of the E, and A the number of nucleons within the effective surface (E) approximation [17, 18, 19, 2, 21]. A separation of the nuclear energy into the volume and surface (and curvature) components of the LDM and ETF makes obviously sense for a/r 1. The accuracy of the E approximation in the ETF approach [11] was checked [19] with the spin-orbit (O) and asymmetry terms [2, 21] by comparing results with those of the Hartree-Fock (HF) and other e-mail: magner@kinr.kiev.ua ETF theories for some kyrme forces. olutions for the isoscalar and isovector particle densities in the E approximation of the ETF approach were applied to the analytical calculations of the surface symmetry-energy constants and the neutron skin in the leading order of the parameter a/r [2, 21]. Our results are compared with the previous ones [1, 2, 3, 4] in the LDM and recent works [5, 6, 7, 8, 22, 23, 24, 25, 26]. The structure of the isovector dipole-resonance (IVDR) strength in terms of the main and satellite (pygmy) modes [23, 24, 25, 26, 27, 28, 29, 3, 31, 32, 33] as a function of the isovector surface energy constant in the E approach can be described within the Fermi-liquid droplet (FLD) model [34, 35]. The analytical expressions for the surface symmetry energy constants are tested by the energies and sum rules of the isovector dipole resonance (IVDR) strength structure within the FLD model [33]. In the present work, we shall extend the variational effective surface method [21, 33] taking into account also derivatives of the non-gradient terms in the symmetry energy density along with its gradient ones. The fundamental isovector derivative and surface-tension constants are not fixed yet well enough by using the experimental data for the neutron skin thickness [8]. We suggest to use also the empirical data for the splitting of the IVDR strength structure to evaluate them better through their analytical ETF expressions in the E approximation as functions of the kyrme force parameters. 2. ymmetry energy and particle densities We start with the nuclear energy as a functional of the isoscalar and isovector particle densities ρ ± = ρ n ± ρ p in the local density approach [11, 13]: E = dr ρ+ E (ρ +,ρ ), where E (ρ +,ρ ) b V +JI 2 +ε + (ρ +,ǫ)+ε (ρ +,ρ,ǫ)+ + (C + /ρ + +D + )( ρ + ) 2 +(C /ρ + +D )( ρ ) 2 (1) 1

and ε (ρ +,ρ,ǫ) = sym (ǫ)(ρ /ρ + ) 2 JI 2, sym (ǫ) = J Lǫ+K ǫ 2 /2. (2) We introduced also a small parameter ǫ of the expansion following suggestions of Myers and wiatecki [1], ǫ = (ρ ρ + )/(3ρ ), where ρ = 3/4πr 3.16 fm 3 is the density of the infinite nuclear matter and r = R/A 1/3 is the radius constant. In (1), b V 16 MeV is the separation energy per particle. The isoscalar part of the surface energy-density(1)(zero-order terms in expansion over ρ /ρ +, denoted as δ [1]) which does not depend explicitly on the density gradient terms, is determined by the function ε + (ρ +,ǫ). We use a representation ε + = K + ε + /18 by the dimensionless quantity ε +. K + is the isoscalar incompressibility modulus of the symmetric nuclear matter, K + 22 245 MeV, ǫ + (ρ +,ǫ) = 9ǫ 2 +I 2 [ sym (ǫ) J]/K +, (3) I = (N Z)/A is the asymmetry parameter, N = drρn (r) and Z = drρ p (r) are the neutron and proton numbers and A = N + Z. The second term, I 2, appears due to the derivative corrections to the volume symmetry energy. The symmetry energy constant of the nuclear matter J 3 MeV specifies the main volume term in the symmetry energy. In contrast to the simplest approximation [21], terms with L 2 12 MeV, and even less known K 47 14 MeV [5, 8], defined by the first and second derivatives of the symmetry energy expansion with respect to the variable ǫ were taken into account(2), (3). Equation(1) can be applied in a semiclassical approximation for realistic kyrme forces [14, 15, 16], in particular by neglecting higher h corrections (the ETF kinetic energy) [11] and Coulomb terms [19, 2, 21]. Up to a small Coulomb exchange terms they all can be easily taken into account [17, 21]. Constants C ± and D ± are defined by parameters of the kyrme forces [11, 13, 14, 15, 16]. For C ± one has ( t 1 25 ), C + = 1 12 C = 1 48 t 1 12 t 2 5 3 t 2x 2 ) ( 1+ 5 2 x 1 1 36 t 2 ( 1+ 19 ) 8 x 2. (4) The isoscalar O gradient terms in (1) are defined with a constant: D + = 9mW 2 /16 h2, where W 1-13 MeV fm 5 and m is the nucleon mass [11, 13, 14, 15]. The isovector O constant D is usually relatively small and will be neglected here for simplicity. We emphasize that (1) has a general form [18, 19, 2] for any densed system with a sharp edge. In particular, it can be derived from comparison with one of energy density functionals determined by a kyrme force in order to get relations (4) for C ± and other ones in terms of its parameters. Equation (1) contains the volume component given by the first two terms and the surface part including the L and K derivative corrections of the ε and density gradients which both are important for the finite nuclear systems [21]. These gradient terms together with other surface terms of the energy density in the E approximation are responsible for the surface tension in finite nuclei. Minimizing the energy E under constraints of the fixed particle number A = dr ρ + (r) and neutron excess N Z = dr ρ (r) (also other constraints as a deformation, nuclear angularmomentum and so on [17, 18, 19, 21]), one arrives at the Lagrange equations with the isoscalar and isovector chemical potentials as corresponding multipliers. The analytical solutions will be obtained approximately up to the order A 2/3 in the binding energy. To satisfy these constraints for calculations of the particle densities, one needs the leading order terms in a/r A 1/3 for calculations of the particle densities ρ ±. Using these densities for the nuclear energy calculations with a required accuracy we account for higher (next) order surface corrections in a/r with respect to the leading order terms [21]. 3. Extended densities and energies For the isoscalar particle density, w = w + = ρ + /ρ, one hasuptotheleading termsin theparametera/rtheusual first-order differential Lagrange equation [19, 2, 21] but with the solution depending on L and K through ε + (3), w x = dy w r 1+βy yǫ + (y,ǫ), x = ξ a, (5) for x < x(w = ) and w = for x x(w = ) where x(w = ) is the turning point, ξ is the distance from a given spatial point to the E in a local coordinate system (ξ,η) (ξ = r R for spherical nuclei and η is the tangentto-e coordinate [18]). ǫ + (y,ǫ) (3) is the dimensionless energy density ε + per particle (1) and β = D + ρ /C + is the dimensionless O parameter. For w r = w(x = ) up to I 2 corrections, one has the boundary condition, d 2 w(x)/dx 2 = at the E (x = ): ǫ + (w r )+w r (1+βw r )[dǫ + (w)/dw] w=wr =. (6) In Eq. (5), a.5.6 fm is the diffuseness parameter [21], C + ρ K + a = 3b 2, (7) V For the isovectordensity, w (x) = ρ /(ρ I), after simple transformations of the isovector Lagrange equation up to the leading term in a/r in the E approximation, one similarly finds the equation and boundary condition for w (w) [21], dw dw = c sym sym (ǫ)(1+βw) 1 w 2, (8) Jǫ + (w) w 2 and w (w = 1) = 1; sym (ǫ) is given by (2) and c sym = a[j/(ρ C )] 1/2, (9) 2

1 Ly5* 1.8 w-.6 - w.1 km* Ly5.4.2 w - w + L=1 MeV L=5 L=.1 Ly5* Ly6 Vsym28 Vsym32-5 -4-3 -2-1 1 2 x=ξ/a Fig. 1: Isovector w (11) [with (L = 5 and 1 MeV) and without (L = ) derivative constant L] and isoscalar w = w + (see[21])particledensitiesareshownvsx = ξ/aforthekyrme force Ly5 (x (r R)/a for small nuclear deformations [16, 33]). see [21] for the case of the constant sym = J. The analytical solution w = wcos[ψ(w)] can be obtained through the expansion of ψ in powers of γ(w) = 3ǫ/c sym. (1) Expanding up to the second order in γ, one finds [36] where w = w cosψ(w) w ( 1 ψ 2 (w)/2+... ), (11) ψ(w) = γ(w)[1+ cγ(w)+...]/ 1+β, (12) c = [βc 2 sym +2+c2 syml(1+β)]/[3(1+β)]. (13) The L dependence of w appears at the second order in γ but w is independent of K at this order. Note that K is the coefficient in expansions (2) and (3) at higher order ǫ 2, and therefore, it shows up at higher (third) order terms of the expansion (11) in γ [21]. InFig.1,theLdependenceofthefunctionw (x)within rather a large interval L = 1 MeV [6] is shown as compared to that of the density w(x) for the Ly5 force asatypicalexample(l = 45.885 5MeV[37]and Table I). As shown in Fig. 2 in a logarithmic scale, one observes a big difference in the isovector densities w derived from different kyrme forces [14, 15, 16] within the diffuse edge. All these calculations have been done with the finite value of the slope parameter L taken approximately from [15, 37] (Table I), which is important for calculations of the neutron skins of nuclei. The isovector particle density w (11) in the second order of the small parameter γ does not depend on the symmetry energy in-compressibility K as mentioned above. Therefore, it is possible to study first the main slope effects of L neglecting small I 2 corrections to the isoscalar particle density w + [21] through the ε + (3). Then, we may deal with more precisely the effect of the second derivatives K taking into account higher order terms..1-1 -.8 -.6 -.4 -.2.2.4.6.8 1 1.2 x=ξ /a Fig. 2: Isovector density w (x) (11) (in the logarithmic scale) as function of x within the quadratic approximation to ε +(w) for several kyrme forces [14, 15, 16]; (L = 48.269 MeV for Ly5 and L = 47.449 MeV for Ly6 [37], which are approximately 5 MeV within a precision of the line thickness; and a similar approximation was used for other forces. We emphasize that the dimensionless densities, w(x) [21] and w (x) (11), shown in Figs. 1 and 2 were obtained in the leading E approximation (a/r 1) as functions of the specific combinations of the kyrme force parameters like β, c sym [21] but taking into account now the L-dependence. They are the universal distributions independent of the specific properties of the nucleus [21]. It yields approximately the local density distributions in the normal-to-e direction ξ with the correct asymptotical behavior outside of the E layer for any deformation at a/r 1, as well as for the semi-infinite nuclear matter. Densities w ± are universal distributions independent of the specific properties of nuclei. The nuclear energy E in this improved E approximation is split into the volume and surface terms [21], E b V A+J(N Z) 2 /A+E (+) +E ( ). (14) For the isoscalar(+) and isovector (-) surface energy components E (±), one obtains E(±) = b (±) /(4πr2 ), where is the surface area of the E, b (±) are the isoscalar (+) and isovector ( ) surface energy constants, b (±) 8πr 2 C ± dξ (1+D ± ρ + /C ± )( ρ ± / ξ) 2. (15) These constants are proportional to the corresponding surface tension coefficients σ ± = b (±) /(4πr2 ) through solutions for ρ ± (ξ) [see (5) and 11)] which can be taken into account in the leading order of a/r. These coefficients σ ± are the same as found in expressions for the capillary pressures of the macroscopic boundary conditions [see [21] extended to new ε ± (2) and (3) modified by L and K derivative corrections]. Within the improved E approximation where higher order corrections in the small parameter a/r are taken into account, we derived equations 3

for the nuclear surface itself. For more exact isoscalar and isovector particle densities we account for the main terms in the next order of the parameter a/r in the Lagrange equations [36]. Multiplying these equations by ρ / ξ and integrating them over the E in the direction ξ and using solutions for w ± (x) up to the leading orders [see (5) and (11)], one arrives at the E equations in the form of the macroscopic boundary conditions [21, 35]. They ensure equilibrium through the equivalence of the volume and surface (capillary) pressure (isoscalar or isovector) variations. The latter ones are proportional to the surface tension coefficients σ ±. For the energy surface coefficients b (±) (15), one obtains b (+) =6C ρ J + + r a, J += and where J = + 1 dw[w(1+βw)ε + (w)] 1/2, (16) b ( ) = k I 2, k = 6C ρ J r a, (17) 1 [ ] 1/2 wǫ+ (w) dw {cosψ (1+βw) [ wsinψ/ (c sym 1+β )] [1+2 cγ(w)]} 2.(18) For γ and c, see (1) and (13). For J one can use the following approximation: 1 { w J (1 w)dw 1+ 2γ(w) 1+βw c sym (1+β) + γ2 (1+β) 2 [ 1 c 2 sym +6(1+β) ( c c sym 1 2 )]}.(19) imple expressions for constants b (±) can be easily derived in terms of the algebraic and trigonometric functions for the quadratic form of ǫ + (w,ǫ) (ǫ = ). In these derivations we neglected curvature terms and, being of the same order, shell corrections. The isovector energy terms were obtained within the E approximation with the high accuracy up to a small I 2 (a/r) 2. According to the macroscopic theory [1, 17, 18, 19, 2, 21], one may define the isovector stiffness Q with respect to the difference R n R p between the neutron and proton radii as a collective variable, Q = k I 2 /τ 2, (2) where τ is the neutron skin in r units. Defining the neutron and proton radii R n,p as the positions of the maxima of the neutron and proton density gradients, respectively, one obtains the neutron skin τ [21], with τ = R n R p 8aIg(w r) r r c 2, (21) sym w 3/2 (1+βw) 5/2 { g(w) = w(1+2 cγ) 2 + (1+β)(3w+1+4βw) 2γ(1+ cγ)[ cw c sym (1+2 cγ)]}, (22) Ly5 Ly5 Vsym28 Vsym32 L(MeV) 5 5 1 6 k, (MeV) -12.6-13.1 11.4 15.6 k (MeV) -13.8-15. 11.6 17.6 ν.37.92.9.84 ν.66.6.83.69 Q (MeV) 73 72-62 -55 Q(MeV) 49 41-56 -37 τ /I.54.43.43.53 τ/i.53.6.45.69 D (MeV) 11 89 78 79 D(MeV) 1 89 78 78 Table I. Isovector energy coefficients k and stiffness Q (23) (in units MeV), factor ν, neutron skin τ/i, and constant D = hω FLD A 1/3 of the IVGDR excitation energy hω FLD (in MeV) for the 132 n are shown for a few kyrme forces at different slope parameters L [22, 15, 37]; the zero low index means that L is neglected; the relaxation time T [4] is the same as used in Figures. taken at the E value w r [w (w r ) = (6)]. Accounting also for (17) and (19), one finally arrives at Q = ν J 2 /k, ν = 9J 2 /[16g2 (w r )], (23) where J and g(w) are given by (17), (19) and (22), respectively. This Q with ν = 9/4 has been predicted earlier [1]. However,inourderivationsν deviatesfrom9/4,andit is proportional to the function J 2 /g2 (w r ). This function depends significantly on the O interaction parameter β but not too much on the specific kyrme forces [21, 33]. The isovector stiffness coefficient Q(23) depend essentially on constants C and L (also K ) through ν (23) and k (17) (and equations (3) and (2)). The universal functions w(x) [21] and w (x) (11) in the leading order of the E approximation can be used [analytically in the quadratic approximation for ε + (w)] for calculations of the surface energy coefficients b (±) (15) and the neutron skin τ I. As it was shown [21], only these particledensity distributionsw(x) and w (x) within the surface layer are needed through their derivatives. Therefore, the surface symmetry-energy coefficient k (17), the neutron skin τ (21) and the isovector stiffness Q (23) can be approximated analytically in terms of functions of the critical surface combinations of the kyrme parameters a, β, b (±), c sym, as well as the volume ones ρ, K, J, and derivatives of the symmetry energy L and K. Thus, they are independent of the specific properties of the nucleus in the E approximation. 4. Discussion of the results In Table I we show the isovector energy coefficient k [Eq. (17)], the isovector stiffness parameter Q (23), its 4

kyrmes L E 1 E 2 1 2 D k Q τ/i MeV MeV MeV % % MeV MeV MeV Ly5 132 n 17.15 19.85 87.9 12.1 88.9-13.1 72.43 5 17.18 19.86 88.8 11.2 88.8-14.2 48.54 6 17.18 19.86 89. 11. 88.8-14.5 45.57 7 17.18 19.87 89.2 1.8 88.7-14.8 42.59 8 17.19 19.87 89.5 1.5 88.8-15.1 4.62 9 17.19 19.87 89.7 1.3 88.8-15.5 38.64 1 17.2 19.87 89.9 1.1 88.8-15.8 36.67 28 Pb 5 14.65 17.14 91.5 8.5 87.9-14.2 48.54 Vsym32 132 n 14.7 18. 72.9 27.1 78.8 15.6-55.53 5 14.69 17.99 75.1 24.9 78.4 17.2-39.66 6 14.69 17.98 75.5 24.5 78.3 17.6-37.69 7 14.69 17.98 76. 24. 78.2 18. -35.71 8 14.69 17.98 76.4 23.6 78.2 18.4-33.74 9 14.68 17.97 76.9 23.1 78.3 18.9-32.77 1 14.68 17.97 77.4 22.6 78. 19.3-3.8 28 Pb 6 12.71 15.56 83.1 16.9 77.7 17.6-37.69 Table II. Energies E n (n = 1,2), EWRs n, average IVDR (IVGDR) energy constants D, surface symmetry energy k and isovector stiffness Q constants, and neutron skin τ/i vs the slope parameter L in a typical region [8] for the same kyrme forces Ly5 or Vsym32 and relaxation time T as in Tables I and Figures; index 1 corresponds to the main mode and index 2 to the satellite one; the results for 28 Pb are shown for the comparison. constant ν and the neutron skin τ (21). They were obtained for a few kyrme forces [14, 15, 16] with different values of L within the E approximation using the quadratic approximation for ε + (w,ǫ = ) [21], and neglecting the I 2 slope corrections. We also show the quantities k,, ν, Q and τ where the slope corrections are neglected (L = ). In contrast to a fairly good agreement for the analytical isoscalar energy constant b (+) [21] with that of [21, 14, 15, 16], the isovector energy coefficients Q (or k = νj 2 /Q) are more sensitive to the choice of the kyrme forces than the isoscalar ones [21]. The modula of Q are significantly smaller for most of the kyrme forces Ly [14, 16] and V [15] than for the other ones, in contrast to the symmetry energy constant k for which one has the opposite behavior. However, the L dependence of k is not more pronounced than that of Q (Table I and II). For Ly and V forces the stiffnesses Q are correspondingly significantly smaller in absolute value being more close to the well-known empirical values of Q [4] and semi-microscopic HF calculations [6, 11], especially with increasing L. Note that the Q is rather more decreased with L than k, sometimes for Ly and V in factor of about two. Thus, the isovector stiffness Q is even much more sensitive to constants of the kyrme force and to the slope parameter L than constants k. wiatecki and his collaborators [4] suggested the isovector stiffness values Q 3 35 MeV accounting for the additional experimental data. More precise A-dependence of the averaged IVDR (Isovector Giant Dipole Resonance, IVGDR) energy D FLD for the finite values of Q seems to be beyond the accuracy of the both hydrodynamical and FLD model calculations. More realistic selfconsistent HF calculations with the Coulomb interaction, surface-curvature and quantum-shell effects lead to larger Q 3 8 MeV [6, 11]. The IVGDR energies and the energy weighted sum rules (EWR) obtained within the semiclassical FLD approach based on the Landau-Vlasov equation [35, 33] with the macroscopic boundary conditions [21] are also basically insensitive to the isovector surface energy constants k or Q. They are similarly in good agreement with the experimental data, and do not depend much on the kyrme forces, also when the finite slope symmetry energy parameter L is included (the two last rows in Table I and 7th column of Table II). An investigation of the splitting of the IVDR within this approach into the main peak which exhausts mainly the model-independent EWR and a much broader satellite (pygmy-like resonances with a much smaller contribution to the EWR [23, 24, 25, 26, 33] ) is extended now by taking into account the slope L dependence of the symmetry energy density. Note that the relative strength of the satellite mode with respect to that of the main peak disappears as I in the limit of symmetric nuclei [34, 35]. The total IVDR strength function being the sum of the two ( out-of-phase n = 1 and in-phase n = 2) modes for the isovector- and isoscalar-like volume particle density vibrations, respectively [Fig. 3 for the finite L = 5 MeV] have a shape asymmetry [33] (the Ly5 forces are taken againasatypicalexample). TheLdependenceofthetotal IVDR strength is shown in Fig. 4 for the Ly5 and Fig. 5 5

.25 132 n Ly5* (ω).2.15.1 tot out in.5 12 14 16 18 2 22 24 26 28 hω, MeV Fig. 3: IVDR strength functions (ω) vs the excitation energy hω are shown for vibrations of the nucleus 132 n for the kyrme force Ly5 by solid line at L = 5MeV; dotted ( out-of-phase ), and dashed ( in-phase ) curves show separately the main and satellite excitation modes, respectively [33]; the relaxation time of the collision term T = 4.3 1 21 s in agreement with widths of the IVGDRs..3 132 n Ly5* (ω).2.1 L=1 MeV L=6 L=5 L= 15 16 17 18 19 2 21 22 23 hω, MeV Fig. 4: The same total strenghts as in Fig. 3 are shown for different L =,5,6 and 1 MeV by dashed, thick and thin solid and dotted lines. 6

.1 132 n Vsym32 (ω).8.6.4 L=1 MeV L=6 L=5 L=.2 13 14 15 16 17 18 19 2 21 hω, MeV Fig. 5: The same as in Fig. 4 but for Vsym32 kyrme force (L = 6 MeV) at the relaxation time T with the same constant of the frequency dependence as in the previous Figure ([4]), T = 7.5 1 21 s. for Vsym32 cases. The characteristic values of L are used in these calculations (L 5 MeV for the Ly5 force [37] and L 6 MeV for the Vsym32 one [15]). In Figs. 3 and 4 for the Ly5 and Fig. 5 for Vsym32 forces, one has the in-phase satellite to the right of the main outof-phase peak. An enhancement on its left for Ly5 is due to the increasing of the out-of-phase strength (dotted) curve at small energies as compared to the IVGDR (Fig. 3 and 4), in contrast to the Vsym32 case shown in Fig. 5. The IVDR energies of the two modes in the nucleus 132 n do not change significantly with L in both cases (Table II). However, as seen from Table II and Figs. 4 and 5, the L dependence of the EWRs of these modes for Vsym32 is larger than for Ly5. Notice, the essential re-distribution of the EWR contributions among them due to a significant enhancement of the main outof-phase peak with increasing L for Ly5 (Figs. 3 and 4) and more pronounced EWR re-distribution with L for Vsym32 (Fig. 5) in the same nucleus (Table II). lopes of the L dependence of τ/i which is almost linear are approximately the same for both considered kyrmes but smaller than those found in [8] probably due to different definitions of the neutron skin thickness τ [21] which are related to the value of δ = ρ /ρ + at the nuclear effective surface [1]. Note also that these more precise calculations (cf. withthoseof[21,33])takeintoaccounthigher(4thorder) terms of the power expansion in the small parameter γ for any reasonable L change [8]. This is essentially important for the IVDR strength distribution for V forces because of smaller c sym as compared to those for other kyrme interactions. Constants c for the isovector solutions w, (11), are modified essentially (besides of the L dependence) by higher order terms due to a non-linearity equation for ψ(w) solved in terms of the power series. Decreasing the relaxation time T in factor of about 1.5 with respect to the value (Figures) evaluated from the data on the widths of the IVGDRs at their energies almost does not change the IVDR strength structure [4]. However, we found its strong dependence on T in a more wide value region [in factor of about 2-3]. The in-phase strength component with rather a wide maximum is weakly dependent on the choice of the kyrme forces [14, 15, 16] and on the slope parameter L, as well as the relaxation time T. The most responsible parameter of the kyrme HF approach leading to the significant differences in the k and Q values is the constant C (4) in the gradient terms of the energy density (1). Indeed, the key quantity in the expression for Q (23) and the isovector surface energy constant k (17), is the constant C because one mainly has k C [21], and Q 1/k 1/C. Concerning k and the IVDR strength structure, this is even more important than the L dependence though the latter changes significantly the isovector stiffness Q and the neutron skin τ. The constant C is very different for different kyrme forces in the absolute value and even sign, C 23 to 26 MeV fm 5 (k 15to18MeV). Contraryto theisoscalar parameters (b (+) ), there are so far no clear experiments which would determine k well enough because the mean energies of the IVGDR (main peaks) do not depend very much on k for different kyrme forces (see last row of Table I and 7th column of Table II). Perhaps, the low-lying isovector collective states are more sensitive but there is 7

no careful systematic study of their k dependence at the present time. Another reason for so different k and Q values might be traced back to the difficulties in deducing k directly from the HF calculations due to the curvature and quantum effects in contrast to b (+). It is worthwhile to study the semi-infinite matter within our approach to avoid such effects and compare with semimicroscopic models. We have to go also far away from the nuclear stability line to subtract uniquely the coefficient k in the dependence of b ( ) I 2 = (N Z) 2 /A 2, according to (17). For exotic nuclei one has more problems to relate k to the experimental data with a good enough precision. The L dependence of the neutron skin τ is essential but not dramatic in the case of Ly and V forces (Table I). The precision of such a description depends more on the specific nuclear models [6]. Our results for the neutron skin R n R p.1.13 fm in 28 Pb (Table II) in a reasonable agreement with the experimental data (Fig. 3 of [8]), (R n R p ) exp =.12.14 fm (the coefficient (3/5) 1/2 of the square-mean neutron and proton radii was taken into account to adopt the definitions, see also very precised experimental data [41]). For 132 n our analytical evaluation predict the neutron skin thickness (R n R p ) exp.12.15 fm (.8.1 fm for a more known nucleus 12 n, also in agreement with the experimental data collected in [8]). The neutron skin thickness τ, like the stiffness Q, is interesting in many aspects for the investigation of exotic nuclei, in particular, in nuclear astrophysics. We emphasize that for the specific kyrme forces there exist an abnormal behavior of the isovector surface constants k and Q as related mainly to the fundamental constant C of the energy density (1) but not much to the derivative symmetry-energy density corrections. The coefficient ν (23) is almost independent on C for Ly and V kyrme forces(table I). As compared to 9/4 suggested in [1], they are significantly smaller in magnitude for the most of the kyrme forces, besides of km* (ν 2.3). Notice that the isovector gradient terms which are important for the consistent derivations within the E approach are not included into the relativistic local density approaches[38, 39]. In contrast to all other kyrme forces, for RATP [14] and V [15] (like for ki) forces, the isovector stiffness Q is even negative as C > (k > ), that would correspond to the unstable vibration of the neutron skin. 5. Conclusions imple expressions for the isovector particle densities and energies in the leading E approximation were obtained analytically, i.e., for the surface symmetry energy, the neutron skin thickness and the isovector stiffness coefficients as functions of the slope parameter L. We have to include higher orderterms in the parametera/r to derivethe surface symmetry energy. It depends on the particle density which can be taken into account at leading order in a/r. These terms depend on the well-known parameters of the kyrme forces. Our results for the isovector surface energy constant k, the neutron skin thickness τ and the stiffness Q depend in a sensitive way on the choice of the parameters of the kyrme energy density(1), especially on its gradient terms through the parameter C (4). Values of the isovector constants k, τ, and especially, Q depend also essentially on the slope parameter L, and the spin-orbit interaction constant β. The isovector stiffness constants Q become more close to the empirical data accounting for their L-dependence. The mean IVGDR energies and sum rules calculated in the macroscopic models like the FLD model [35, 33] are in a fairly good agreement with the experimental data. We found a reasonable two-mode main and satellite structure of the IVDR strength within the FLD model as compared to the experimental data and recent other theoretical works. We may interprete semiclassically the IVDR satellites as some kind of the pygmy resonances. Their energies and sum rules obtained analytically within the semiclassical FLD approximation are sensitive to the surface symmetry energy constant k. Energies E 1 and E 2 (Table II) are independent of the slope parameter L but EWRs 1 and 2 do depend on L, especially for Vsym32. It seems helpful to describe them in terms of the only few critical constants, like k and L. Therefore, their comparison with the experimental data on the IVDR strength splitting can be used for the evaluation of k and L, in addition to the experimental data for the neutron skin. For further perspectives, it would be worth to apply our results to calculations of pygmy resonances in the IVDR strength within the FLD model [35] in a more systematic way. More general problems of the classical and quantum chaos in terms of the level statistics [42] and Poincare and Lyapunov exponents [43, 44] might lead to a progress in studying the fundamental properties of the collective dynamics like nuclear fission within the wiatecki& trutinsky Macroscopic-Microscopic model. Our approach is helpful also for further study of the effects in the surface symmetry energy because it gives the analytical universal expressions for the constants k, τ and Q as functions of the symmetry slope parameter L which are independent of the specific properties of the nucleus. Acknowledgement Authors thank K. Arita, N. Dinh Dang, M. Kowal, V.O. Nesterenko, M. Matsuo, K. Matsuyanagi, J. Meyer, T. Nakatsukasa, A. Pastore, P.-G. Reinhard, A.I. anzhur, J. kalski, and X. Vinas for many useful discussions. One of us (A.G.M.) is also very grateful for a nice hospitality during his working visits at the National Centre for Nuclear Research in Otwock-wierk and Warsaw, Poland, and at the Nagoya Institute of Technology. This work was partially supported by the Deutsche Forschungsgemeinschaft Cluster of Excellence Origin and tructure of the Universe (www.universe-cluster.de) and Japanese ociety of Promotion of ciences, ID No. -1413. 8

References [1] Myers W D, wiatecki W J 1969 Ann. Phys. 55 395; 1974 84 186 [2] Myers W D, wiatecki W J 198 Nucl. Phys. A336 267 [3] Myers W D, wiatecki W J and Wang C 1985 Nucl. Phys. A436 185 [4] Myers W D, wiatecki W J 1996 Phys. Rev. C61 141 [5] Centelles M, Roca-Maza X, Vinas X, and Warda M 29 Phys. Rev. Lett. 12 1252 [6] Warda M, Vinas X, Roca-Maza X and Centelles M 21 Phys. Rev. C 82 54314 [7] Roca-Maza X, Centelles M, Vinas X and Warda M 211 Phys. Rev. Lett. 16 25251 [8] Vinas X, Centelles M, Roca-Maza X and Warda M 214 Eur. Phys. A 5 27 [9] Runge E and Gross E K U 1984 Phys. Rev. Lett. 52 997 [1] Marques M A L and Gross E K U 24 Annu. Rev. Chem., 55 427 [11] Brack M, Guet C, and Hakansson H-B 1985 Phys. Rep. 123 275 [12] Vautherin D and Brink D M 1972Phys. Rev. C 5 626 [13] Bender M, Heen P-H, and Reinhard P-G 23 Rev. Mod. Phys. 75 125 [14] Chabanat E et al 1997 Nucl.Phys. A627 71; 1998 A635 231 [15] Klüpfel P, Reinhard R-G, Bürvenich T J and Maruhn J A 29 Phys. Rev. C79 3431 [16] A. Pastore et al 213 Phys. cr. T154 1414 [17] trutinsky V M and Tyapin A 1964 Exp. Theor. Phys. (UR) 18 664 [18] trutinsky V M, Magner A G and Brack M 1984 Z. Phys., A 319 25 [19] trutinsky V M, MagnerA G and Denisov V Yu 1985 Z. Phys. A 322 149 [2] Magner A G, anzhur A I and Gzhebinsky A M 29 Int. J. Mod. Phys. E 18 885 [21] Blocki J P, Magner A G, Ring P and Vlasenko A A 213 Phys. Rev. C87 4434 [22] AgrawalB K, De J N, amaddarkat al 213Phys. Rev. C 87 5136 [23] Voinov A et al 21 Phys. Rev. C81 24319 [24] Larsen A C et al 213, Phys. Rev. C87 14319 [25] Repko A, Reinhard R-G, Nesterenko V O and Kvasil J 213 Phys. Rev. C87 2435 [26] Kvasil J, Repko A, Nesterenko V O et al 213 Phys. cr. T154 1419 [27] Dinh Dang N, Tanabe K and Arima A 1998 Phys. Rev. C 58 3374; 1999 Phys. Rev. C 59 3128 [28] Dinh Dang N, Kim Au V, uzuki T and Arima A 21 Phys. Rev. C 63 4432 9 [29] Ryezayeva N, Hartmann T, Kalmykov Y et al 22 Phys. Rev. Lett. 89 27252 [3] Adrich P et al 25 Phys. Rev. Lett. 95 13251 [31] Wieland O et al 29 Phys. Rev. Lett. 12 9252 [32] Tohyama M and Nakatsukasa T 212 Phys. Rev. C 85 3132(R) [33] Blocki J P, Magner A G and Ring P 214 Phys. cr. T89 5419 [34] Kolomietz V M and Magner A G 2 Phys. Atom. Nucl. 63 1732 [35] Kolomietz V M, Magner A G, and hlomo 26 Phys. Rev. C73 24312 [36] Blocki J P, Magner A G and Ring P 215 to be published. [37] Meyer J 214, private communications. [38] Vretenar D, Paar N, Ring P and Lalazissis G A 21 Phys. Rev. C63 R13 [39] Vretenar D, Paar N, Ring P and Lalazissis G A 21 Nucl. Phys. A692 496 [4] Magner A.G, Gorpinchenko D V and Bartel J 214 Phys. At. Nucl. 77 1229 [41] Tamii et al 211 Phys. Rev. Lett. 17 6252 [42] Blocki J P, Magner A G and Yatsyshyn I 21 At. Nucl. Energy 11 239 [43] Blocki J P, Magner A G and Yatsyshyn I 211 Int. J. Mod. Phys. E 2, 292 [44] Blocki J P and Magner A G 212 Phys. Rev. C 85 64311