Statistical Approach to Nuclear Level Density

Similar documents
Physics Letters B 702 (2011) Contents lists available at ScienceDirect. Physics Letters B.

arxiv: v1 [nucl-th] 4 Feb 2011

New T=1 effective interactions for the f 5/2 p 3/2 p 1/2 g 9/2 model space: Implications for valence-mirror symmetry and seniority isomers

Quantum Chaos as a Practical Tool in Many-Body Physics ESQGP Shuryak fest

Quantum Chaos as a Practical Tool in Many-Body Physics

Correlations between magnetic moments and β decays of mirror nuclei

Spectral Fluctuations in A=32 Nuclei Using the Framework of the Nuclear Shell Model

arxiv: v1 [nucl-th] 5 Apr 2017

Projected shell model for nuclear structure and weak interaction rates

Level density for reaction cross section calculations

The Nuclear Many-Body Problem. Lecture 2

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

ISSN : Asian Journal of Engineering and Technology Innovation 02 (03) 2014 (08-13) QR Code for Mobile users

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

Moment (and Fermi gas) methods for modeling nuclear state densities

Nuclear Structure Theory II

Interaction cross sections for light neutron-rich nuclei

Medium polarization effects and pairing interaction in finite nuclei

Probing the evolution of shell structure with in-beam spectroscopy

Features of nuclear many-body dynamics: from pairing to clustering

Novel High Performance Computational Aspects of the Shell Model Approach for Medium Nuclei

Accurate Description of the Spinand Parity-Dependent Nuclear Level Densities

Nuclear Level Density with Non-zero Angular Momentum

Pairing Interaction in N=Z Nuclei with Half-filled High-j Shell

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

Probing shell evolution with large scale shell model calculations

arxiv:nucl-ex/ v1 29 Apr 1999

Shape Coexistence and Band Termination in Doubly Magic Nucleus 40 Ca

arxiv: v1 [nucl-th] 8 Sep 2011

The Nuclear Many Body Problem Lecture 3

Physics Letters B 695 (2011) Contents lists available at ScienceDirect. Physics Letters B.

Nucleon Pair Approximation to the nuclear Shell Model

Joint ICTP-IAEA Workshop on Nuclear Structure Decay Data: Theory and Evaluation August Introduction to Nuclear Physics - 1

Shell model Monte Carlo level density calculations in the rare-earth region

Probing the shell model using nucleon knockout reactions

Configuration interaction approach to nuclear clustering

Coupled-cluster theory for nuclei

Triune Pairing Revelation

Computational approaches to many-body dynamics of unstable nuclear systems

Configuration interaction studies of pairing and clustering in light nuclei

arxiv:nucl-th/ v1 24 Aug 2005

Renormalization Group Methods for the Nuclear Many-Body Problem

Mihai Horoi, Department of Physics, Central Michigan University, Mount Pleasant, Michigan 48859, USA

arxiv:nucl-th/ v1 5 Aug 1997

arxiv:nucl-th/ v1 10 Jan 2002

Shell model description of dipole strength at low energy

QUANTUM CHAOS IN NUCLEAR PHYSICS

Direct reactions methodologies for use at fragmentation beam energies

shell Advanced Science Research Center, Japan Atomic Energy Agency

The nuclear shell-model: from single-particle motion to collective effects

Exact solution of the nuclear pairing problem

arxiv: v2 [nucl-th] 8 May 2014

APPLICATIONS OF A HARD-CORE BOSE HUBBARD MODEL TO WELL-DEFORMED NUCLEI

Introduction to NUSHELLX and transitions

Extrapolation of neutron-rich isotope cross-sections from projectile fragmentation

Charge density distributions and charge form factors of some even-a p-shell nuclei

RFSS: Lecture 8 Nuclear Force, Structure and Models Part 1 Readings: Nuclear Force Nuclear and Radiochemistry:

Three-nucleon forces and shell structure of neutron-rich Ca isotopes

Generalized seniority scheme in light Sn isotopes. Abstract

Observables predicted by HF theory

Evaluation of inclusive breakup cross sections in reactions induced by weakly-bound nuclei within a three-body model

Aligned neutron-proton pairs in N=Z nuclei

Title. Author(s)Itagaki, N.; Oertzen, W. von; Okabe, S. CitationPhysical Review C, 74: Issue Date Doc URL. Rights.

Nuclear structure input for rp-process rate calculations in the sd shell

Core Polarization Effects on the Inelastic Longitudinal C2 and C4 form Factors of 46,48,50 Ti Nuclei

Spectroscopic overlaps between states in 16 C and 15 C IV (with WBT and NuShell)

The shell model Monte Carlo approach to level densities: recent developments and perspectives

arxiv: v1 [nucl-th] 5 Nov 2018

arxiv: v1 [nucl-th] 7 Nov 2018

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

Nuclear Shell Model. Experimental evidences for the existence of magic numbers;

Thermodynamics of nuclei in thermal contact

c E If photon Mass particle 8-1

Shell evolution and pairing in calcium isotopes with two- and three-body forces

Fitting Function for Experimental Energy Ordered Spectra in Nuclear Continuum Studies

New Trends in the Nuclear Shell Structure O. Sorlin GANIL Caen

A simple effective interaction for 9 He, and Gamow-SRG

Institut d Astronomie et d Astrophysique, Université Libre de Bruxelle, Belgium

On unitarity of the particle-hole dispersive optical model

Spherical-deformed shape coexistence for the pf shell in the nuclear shell model

Comparison of Nuclear Configuration Interaction Calculations and Coupled Cluster Calculations

Statistical properties of nuclei by the shell model Monte Carlo method

Lecture 4: Nuclear Energy Generation

Shell model calculation ー from basics to the latest methods ー

PHYSICAL REVIEW C 74, (2006) (Received 23 January 2006; published 15 September 2006)

Probing neutron-rich isotopes around doubly closed-shell 132 Sn and doubly mid-shell 170 Dy by combined β-γ and isomer spectroscopy.

(Multi-)nucleon transfer in the reactions 16 O, 3 32 S Pb

This is an electronic reprint of the original article. This reprint may differ from the original in pagination and typographic detail.

PHYSICAL REVIEW C 80, (2009) East Lansing, Michigan , USA (Received 12 June 2009; published 1 September 2009)

Structure of Sn isotopes beyond N =82

The structure of neutron deficient Sn isotopes

Lecture 13: The Actual Shell Model Recalling schematic SM Single-particle wavefunctions Basis states Truncation Sketch view of a calculation

Mean field studies of odd mass nuclei and quasiparticle excitations. Luis M. Robledo Universidad Autónoma de Madrid Spain

Nuclear Spectroscopy I

Band crossing and signature splitting in odd mass fp shell nuclei

An improved nuclear mass formula: WS3

Investigation of the Nuclear Structure for Some p-shell Nuclei by Harmonic Oscillator and Woods-Saxon Potentials

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

Part III: The Nuclear Many-Body Problem

Continuum Shell Model

Transcription:

Statistical Approach to Nuclear Level Density R. A. Sen kov,v.g.zelevinsky and M. Horoi Department of Physics, Central Michigan University, Mount Pleasant, MI 889, USA Department of Physics and Astronomy and National Superconducting Cyclotron Laboratory, Michigan State University, East Lansing, MI 88-131,USA Abstract. We discuss the level density in a finite many-body system with strong interaction between the constituents. Our primary object of applications is the atomic nucleus but the same techniques can be applied to other mesoscopic systems. We calculate and compare nuclear level densities for given quantum numbers obtained by different methods, such as nuclear shell model (the most successful microscopic approach), our main instrument - moments method (statistical approach), and Fermigas model; the calculation with the moments method can use any shell-model Hamiltonian excluding the spurious states of the center-of-mass motion. Our goal is to investigate statistical properties of nuclear level density, define its phenomenological parameters, and offer an affordable and reliable way of calculation. Keywords: Nuclear level density, Fermi-gas model, parity, spin PACS: 1.1.Ma, 1.1.Dr, 1.1.Hw, 1..Cs MOMENTS METHOD To describe level densities of a nuclear system in the moments method [1, ] one needs to introduce a restricted single-particle model space with an effective interaction, as it is usually done in the shell model, H = ε +V, (1) where ε and V are the one- and two-body parts of the Hamiltonian, correspondingly. The many-body Hilbert space of the problem is divided into partitions κ that are fixed by the occupation numbers of each single-particle orbit. For example, for the case of 8 Si in the sd model space there are six valence protons and six valence neutrons to be distributed over the three single-particle orbits, d /, s 1/, and d 3/. There are 1 proton and 1 neutron partitions (i.e. 1 different ways of distributing six particles over three sd orbits), which results in total partitions. Each partition can be divided further into smaller subspaces corresponding to exact quantum numbers, such as total spin and parity, α = J π. Let us consider the nuclear level density (NLD) for a given partition κ and a set of quantum numbers α. Sucha partial NLD is not well defined, it explicitly depends on the choice of a single-particle basis (the partition κ is not a good quantum number). However, it is known that this partial NLD in the original harmonic-oscillator basis is close to a Gaussian [3], so that the total NLD can be presented as a sum over all the partitions, ρ(e; α)= D κα G κα (E), () κ where G κα (E) is a finite range Gaussian, D κα is the normalization factor corresponding to the total number of manybody states constructed within a given partition κ for quantum numbers α,ande is the excitation energy. To calculate the finite range Gauss distributions G κα one needs to find the position of each Gaussian (centroid or the first moment), the width (the second moment), and the ground state energy (for more details see, for example, [, ], where also the algorithm was presented for subtracting spurious states related to the motion of the nucleus as a whole). For each configuration (κα) we define the first and the second moments of the Hamiltonian (1) as E κα = H κα and σ κα = H κα H κα. (3) Here the average of an operator O is defined as a normalized sum of all diagonal matrix elements calculated within the given configuration, O κα = 1 D κα Tr (κα) [O]= 1 D κα n O n. () n (κα) The Fourth Conference on Nuclei and Mesoscopic Physics 1 AIP Conf. Proc. 119, 1-18 (1); doi: 1.13/1.8993 1 AIP Publishing LLC 978--73-1-9/$3. 1 This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 3.9..11 On: Mon, Mar 1 ::1

1 J= 3 1 J=1 8 8 J= J=3 8 8 FIGURE 1. Comparison of the shell-model NLD (solid curves) with the moments method densities (dashed curves) calculated for 8 Si, sd model space, J=, 1,, and 3, positive parity. USD effective interaction was used [9]. The interaction term in the Hamiltonian (1) fully accounts for the off-diagonal elements between the partitions. The details of calculation of the traces (3) and (), explanation of the computational algorithm, and the code itself can be found in Ref. []. The last ingredient needed for the NLD calculation is the ground state energy that cannot be determined within the moments method. The ground state energy can be obtained using the standard shell model approach for the same model space and the same effective interaction [] or, for example, using the exponential convergence method [7, 8]. RESULTS Fig. 1 provides an excellent illustration to the moments method. It presents the comparison of the NLD for the 8 Si calculated within the full shell-model diagonalization machinery (solid curves) and by the moments method (dashed curves). Both calculations were performed in the same sd model space and with the same USD effective interaction [9]. Different panels show the NLD corresponding to different spins, J=, 1,, and 3; all of them demonstrate perfect agreement between the shell model and moments method. The agreement confirms the ideas of quantum chaos and complexity of the majority of states in a closed mesoscopic system of strongly interacting particles put in the foundation of the statistical approach [3, 1]. Fig. compares the NLD of 8 Si calculated within the moments method (straight curves) to the experimental level densities (stair curves). There are two stair-case lines on each figure: the solid stair lines present an "optimistic" approach, when all experimental levels with uncertain parity are counted; oppositely, the dashed stair lines present a "pessimistic" approach, when only the experimental levels with assigned parity are included. Again, one can see a fairly good agreement between the densities, of course limited by the space 13 This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 3.9..11 On: Mon, Mar 1 ::1

3 1 π=+1 3 1 π= 1 8 1 1 1 8 1 1 1 FIGURE. 8 Si, all J, both parities. Comparison of nuclear level densities from experiment (stair curves) and calculated with the moments method (straight line). Calculation were performed in a s-p-sd-pf model space, up to 1 hω excitation were included, with WBT effective interaction [17]. truncation. Summarizing the results from Figs. 1 and we can conclude that the moments method is able to provide a reasonably good description of the nuclear level density, at least for not very high excitation energy. The natural next step in development of theory should be an attempt, in application to many nuclei, to extract the global empirical characteristics, such as temperature, entropy, parameters of phenomenological fits, spin cutoff, and so on in order to bridge the gap between the shell-model theory and convenient statistical descriptions traditionally used by experimentalists. The standard statistical approaches to NLD were started long ago [11-1]. In such approaches the nuclear levels are treated as the excited states of a Fermi-gas system characterized by excitation energy E corresponding to temperature T. In a statistical consideration, the NLD at given energy is exponentially proportional to the entropy of the system, ρ exp(s). The more accurate expression for the Fermi-gas NLD can be presented as ρ(e,m)= π 1a 1/ E / 1 πσ exp ( ae M σ ), () where E is the excitation energy, M is the nuclear spin projection, σ is the spin cut-off parameter which is generally related to an effective moment of inertia and temperature, and a is the single-particle level density parameter defined at low energies from E at. The Fermi-gas model formula () is quite attractive due to its simplicity and minimalistic computational demands as compared to the shell-model or other microscopic approaches, such as the HFB combinatorial method [18, 19, ]. In general, one can hope that, with excitation energy growing, we deal with very complicated chaotic wave functions, while the regular statistical features related to the mean field and Fermi-statistics will survive. On the other hand, treating a real nucleus as a low-temperature Fermi-gas is a rough approximation, the distribution over spin projections and parity is usually introduced phenomenologically, and the density parameters such as a and σ have to be fit to individual nuclei. In this work we compare the real NLD (calculated with the moments method in a realistic model space) with the Fermi-gas description. From such a comparison we hope to be able to conclude how good the traditional Fermi-gas model is for a description of realistic nuclear systems and try to find the optimal Fermi-gas parameters and their evolution along the nuclear chart. First we discuss the spin cut-off parameter σ. Following the standard ideas of random angular momentum coupling, we can plot ln[ρ(e,m)] as a function of M ln[ρ(e,m)] = ln[ρ(e,)] 1 σ M, () where ln(ρ(e,)] = ae ln[e]+constant. (7) Fig. 3 presents the corresponding plots for 8 Si, Fe, Ca, and Cr isotopes. Different colors refer to different excitation energies: MeV (black), 1 MeV (red), 1 MeV (green), MeV (blue), and MeV (brown). Dots 1 This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 3.9..11 On: Mon, Mar 1 ::1

Log[ρ(Ε,Μ)] 1 8 Si E= MeV E=1 MeV E=1 MeV E= MeV E= MeV Log[ρ(Ε,Μ)] 1 1 Fe E= MeV E=1 MeV E=1 MeV E= MeV E= MeV - 8 1 8 1 Log[ρ(Ε,Μ)] Ca E= MeV E=1 MeV E=1 MeV E= MeV E= MeV Log[ρ(Ε,Μ)] Cr E= MeV E=1 MeV E=1 MeV E= MeV E= MeV - - - 8 1 M - 8 1 M FIGURE 3. Logarithm of level density, ln[ρ(e,m)], versusm, see Eq. (), calculated for 8 Si, Fe, Ca, and Cr. Different colors present different excitation energies: MeV (black), 1 MeV (red), 1 MeV (green), MeV (blue), and MeV (brown). Dots correspond to the moments-method calculations, the solid lines are linear interpolations, see the text. correspond to the moments-method densities. The solid lines show the best linear interpolations, for each line we can determine the slope that can be related to the spin cut-off parameter and the constant (at M = ) that corresponds to ln[ρ(e,)], see Eq. (). From Fig. 3 we see that for 8 Si and Fe isotopes ln[ρ] is linearly proportional to M with a good accuracy; the slope and the constant values depend very smoothly on the excitation energy. But for the isotopes Ca and Cr the situation is different: the linear approximation is not good and the energy dependence is irregular. It turns out that Ca and Cr belong to a group of nuclei where NLD cannot be well described with the Fermi-gas formula (). In all nuclei of this type, the isospin- channel of interaction is suppressed and only the isospin-1 channel works. Indeed, the Ca nucleus has only four valence neutrons in the pf model space, so that only the isospin-1 part of the two-body effective interaction V from Eq. (1) works. The same happens to Cr, in the pf model space all neutron states (N = ) are occupied, the active degrees of freedom correspond to only four protons which again switches off the isospin- part of the interaction. For the future interpolation procedures we will keep this note in mind. Assuming that the linear approximation () approximately works (at least for a set of nuclei) we can try to extract the spin cut-off parameter σ. Forσ it is natural to suggest the following energy (temperature) dependence, since σ T E: σ = α E(1 + βe), (8) where α and β are the fitting parameters. To fit these parameters we selected the energy interval from to MeV to be used for interpolation. The excitation energy should be sufficiently high to justify the statistical methods and sufficiently low to be able to use a restricted model space. Fig. presents the results of such an interpolation according to Eq. (8). On the left panel one can see α and β parameters fitted for isotopes of calcium (red dots), scandium (green 1 This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 3.9..11 On: Mon, Mar 1 ::1

α-parameter 8 All others Z= isotopes Z=1 isotopes Z= isotopes N= isotones α-parameter 8 Moments Method Statistical calc Rigid sphere 3 3 7 7 8 3 3 7 7 8.. β-parameter -. β-parameter -. 3 3 7 7 8 3 3 7 7 8 FIGURE. Interpolation of the spin cut-off parameter σ, see Eq. (8). Parameters α and β are shown for different sd and pf nuclei. Left panel: different colors present different isotopes or isotones. Right panel: moments-method calculation with interpolation (black circles), statistical calculation (orange diamonds), rigid sphere approximation (yellow squares). dots), titanium (blue dots), and N = isotones (brown dots). The black dots present all the other sd and pf nuclei. The isotopes with the zero ground state isospin (coloured dots) expose a clear shell-structure behavior and are slightly off line. The right panel of Fig. shows the comparison of the α parameter from the moments-method calculations (all black dots) with statistical calculations (σ gt M, orange diamonds) and with the rigid sphere moment-of-inertia approximation (σ TM p A /3, yellow squares). We can conclude that the shell-model results are in a remarkable agreement with both statistical and moment-of-inertia based calculations. It is interesting to note that, inside a separate shell (the left island for the sd shell and the right island for the pf shell), σ goes down opposite to the statistical and rigid sphere predictions which would lead to a growth with mass number A as σ A /3. Fig. presents the similar interpolation results for the level density parameter a, see Eq. (7), that is typically considered as coming from the single-particle level density at the Fermi surface. The colouring scheme on the left panel coincide with the corresponding scheme from Fig.. Again we can clearly see the shell-structure effects and the isospin-suppressed nuclei are slightly off the common picture. The right panel shows the comparison between the moments-method calculations of this parameter (black circles), the fit from experimental data on neutron resonances (Ref. [1], orange diamonds), and the fit using experimental low-lying levels ([], yellow squares). We see that the level density parameter a calculated with the moments method is noticeably lower than the corresponding value extracted from fitting to experimental data. It could mean that the moments method NLD grow up with the excitation energy not as fast as the experimental densities. The situation is not clear and demands additional investigation, we leave it for the near future. A natural explanation of the observed discrepancy is that the moments-method NLD (and the shell-model as such) is loosing some levels at higher excitation energies due to the restricted model space. We have presented the first results of studying the nuclear level density by its (practically exact) calculation in the framework of the shel model with mean-field single-particle orbitals and effective residual interaction that describes rather well the low-lying spectroscopic data. The moments method essentially exactly reproduces the level density for given exact quantum numbers as a function of excitation energy avoiding the diagonalization of prohibitively large Hamiltonian matrices. The justification of the method is in complexity of the many-body wave functions that brings the dynamics into the stage of thermalization and quantum chaos. The validity of the approach is intrinsically limited by energies where the contribution from the orbital space excluded from the shell model becomes important. By our 1 This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 3.9..11 On: Mon, Mar 1 ::1

a-parameter 1 All others Z= Isotopes Z=1 Isotopes Z= Isotopes N= Isotones a-parameter 1 Moments Method FG: Rohr FG: Al-Quraishi 3 7 8 3 7 8 constant constant - 3 7 8-3 7 8 FIGURE. Interpolation of the single-particle level density parameter a, see Eq. (7). Left panel: different colors present different isotopes or isotones. Right panel: moments-method calculation with interpolation (black circles), fit using the experimental data on neutron resonances (Ref. [1], orange diamonds), and fit using experimental low lying levels ([], yellow squares). estimates, it can be used rather safely at least up to 1- MeV excitation energy. The limitation from above comes also from the continuum that converts mean-field levels into the resonances, so that the concept of the discrete level density becomes less definite when the resonances start to overlap. We just started the comparison of the results of the moments method with standard phenomenological expressions used by generations of physicists. There are still many open questions which we hope to address in the future. An interesting physical problem to be discussed elsewhere is the understanding of the role played by various components of the residual interaction in forming the level density. ACKNOWLEDGMENTS Support from the NUCLEI SciDAC Collaboration under U.S. Department of Energy Grant No. DE-SC89 is acknowledged. V.Z. and M.H. also acknowledge U.S. NSF Grants No. PHY-1 and PHY-1817. REFERENCES 1. S. S. M. Wong, Nuclear Statistical Spectroscopy, Oxford, University Press, 198.. V.K.B.KotaandR.U.Haq,Spectral Distributions in Nuclei and Statistical Spectroscopy, World Scientific, Singapore, 1. 3. T.A.Brody,J.Flores,J.B.French,P.A.Mello,A.Pandey,andS.S.M.Wong, Rev. Mod. Phys. 3, 38 (1981).. R. A. Sen kov, M. Horoi, and V. Zelevinsky, Phys. Lett. B 7, 13 (11).. R. A. Sen kov, M. Horoi, and V. Zelevinsky, Comp. Phys. Comm. 18, 1 (13).. NuShellX@MSU, B.A. Brown, W.D.M. Rae, E. McDonald, and M. Horoi, http://www.nscl.msu.edu/~brown/resources/resources.html 7. M. Horoi, A. Volya, and V. Zelevinsky, Phys. Rev. Lett. 8, (1999). 8. M. Horoi, B. A. Brown, and V. Zelevinsky, Phys. Rev. C, 733 (). 9. B. A. Brown and B. H. Wildenthal, Annu. Rev. Nucl. Part. Sci. 38, 9 (1988). 1. V. Zelevinsky, B. A. Brown, N. Frazier, and M. Horoi Phys. Rep. 7, 8 (199). 17 This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 3.9..11 On: Mon, Mar 1 ::1

11. Ya. I. Frenkel, Sow. Phys. 9, 33 (193). 1. H. Bethe, Phys. Rev., 33 (193). 13. H. Bethe, Rev. Mod. Phys. 9, 9 (1937). 1. L. D. Landau, J. Exp. Theor. Phys. 7, 819 (1937). 1. T. Ericson, Adv. Phys. 9, (19). 1. A. Gilbert and A. G. W. Cameron, Can. J. Phys. 3, 1 (19). 17. E. K. Warburton and B. A. Brown, Phys. Rev. C, 93 (199). 18. S. Goriely, S. Hilaire, and A. J. Koning, Phys. Rev. C 78, 37 (8). 19. S. Goriely, S. Hilaire, A. J. Koning, M. Sin, and R. Capote, Phys. Rev. C 79, 1 (9).. http://www.astro.ulb.ac.be/pmwiki/brusslib/level 1. G. Rohr, Z. Phys. A 318, 99 (198).. S. I. Al-Quraishi, S. M. Grimes, T. N. Massey, and D. A. Resler, Phys. Rev.C3, 83 (1). 18 This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 3.9..11 On: Mon, Mar 1 ::1