Natural orbital description of the halo nucleus 6He

Special Section on the Celebration for 55 Years’ Dedicated Research on Heavy-Ion Physics of Natowitz and his 80th Birthday

Natural orbital description of the halo nucleus 6He

Chrysovalantis Constantinou
Mark A. Caprio
James P. Vary
Pieter Maris
Nuclear Science and TechniquesVol.28, No.12Article number 179Published in print 01 Dec 2017Available online 21 Nov 2017
7801

Ab initio calculations of nuclei face the challenge of simultaneously describing strong short-range internucleon correlations and the long-range properties of weakly-bound halo nucleons. Natural orbitals, which diagonalize the one-body density matrix, provide a basis which is better matched to the physical structure of the many-body wave function. We demonstrate that the use of natural orbitals significantly improves convergence for ab initio no-core configuration interaction calculations of the neutron halo nucleus 6He, relative to the traditional oscillator basis.

Neutron halo nucleus 6HeNuclear structureNuclear theory

1 Introduction

Ab initio calculations of nuclear structure [1-9] face the challenge of describing a complex, multiscale quantum many-body system. The goal is to directly solve the many-body problem for a system of protons and neutrons, with realistic internucleon interactions [10-13]. However, the nucleus is governed by a strong, short-range interaction. Short-range correlations, tightly-bound α clusters [14], and weakly-bound halo nucleons [15, 16] introduce dynamics over differing length scales and energy scales, within the same nucleus, which must be simultaneously described within the same many-body calculation.

Natural orbitals [17-22] provide a means of adapting the single-particle basis to better match the physical structure of the many-body wave function. Natural orbitals are obtained by diagonalizing the one-body density matrix, deduced from a preliminary many-body calculation using an initial reference single-particle basis. The Laguerre function basis [23, 24] has been used as the starting point for natural orbitals in atomic electron-structure calculations [18], while we start from the harmonic oscillator orbitals [25] more familiar to the nuclear structure context [26]. The natural orbital basis builds in important contributions from high-lying orbitals of the initial basis — for the present application, high-lying oscillator shells — thereby accelerating the convergence of wave functions, energies, and other observables.

In this letter, we present a framework for ab initio no-core configuration interaction (NCCI) [8] calculations with a natural orbital basis and demonstrate improved convergence for the lightest neutron halo nucleus 6He [15]. When used with recently-proposed infrared (IR) basis-extrapolation schemes [27, 28], we show that natural orbitals provide improved independence of basis parameters for predictions of energy and radius observables.

2 Natural orbitals

In NCCI calculations, the nuclear many-body Schrödinger equation is formulated as a matrix eigenproblem, where the Hamiltonian is represented within a basis of Slater determinants, i.e., antisymmetrized products of single-particle states. Conventionally, harmonic oscillator orbitals [25] are used, and the basis is truncated to a maximum allowed number Nmax of oscillator excitations [8]. The calculated wave functions, energies, and observables depend upon both the truncation Nmax and the oscillator length b of the basis (or, equivalently, the oscillator energy ħwb-2). The solution of the full, untruncated many-body problem could, in principle, be obtained to any desired accuracy, by retaining a sufficiently complete basis set. However, the dimension of the NCCI problem increases rapidly with the number of nucleons and included single-particle excitations, as shown in Fig. 1. Currently available computational resources therefore limit the convergence of calculated states and observables [29-31].

Figure 1:
Growth of the NCCI problem dimension as a function of the number of oscillator excitations Nmax included in the basis, for selected nuclides, including 6He (red curve). The dimensions shown are for spaces with zero angular momentum projection (M=0) and positive parity.
pic

We therefore seek a physically-adapted basis, in which the nuclear many-body wave function can be efficiently and accurately described, subject to the constraint of accessible problem dimensions. The natural orbital basis minimizes the mean occupation of states above the Fermi surface [21, 32], thus reducing the contribution of high-lying orbitals in describing the many-body wave function. Intuitively, the natural orbitals may be understood as representing an attempt to recover the basis in which the many-body wave function most resembles a single Slater determinant. Although we cannot expect to reduce the complex, highly-correlated nuclear wave function to a single Slater determinant, as assumed in the Hartree-Fock approximation, we may expect that transforming to a more natural single-particle basis could enhance the role of a comparatively small set of dominant Slater determinants, and thereby accelerate convergence within a Slater determinant expansion.

An NCCI state of good total angular momentum can only, in general, be obtained as a superposition of several Slater determinants of nljm single particle states (here, n is the radial quantum number, l labels the orbital angular momentum, j labels the resultant angular momentum after coupling to the spin, and m labels its projection). Therefore, we cannot, in general, expect the nuclear eigenfunctions to resemble a single Slater determinant. Rather, we hope to recover wave functions which most resemble a single configuration of nucleons over nlj orbitals. For this purpose, we consider the scalar densities ρab(0)Ψ[cbc˜a]00|Ψ, where ca represents the creation operator for a nucleon in orbital a=(nalaja) and the brackets [...]00 represent spherical tensor coupling to zero angular momentum. For a single configuration |Ψ〉, the scalar densities are diagonal, and the diagonal entries give the occupations of the contributing orbitals [Na=(2ja+1)1/2ρaa(0)] [33]. Otherwise, for the general case of a many-body state |Ψ〉, natural orbitals are defined by diagonalizing this scalar density matrix. The scalar density matrix only connects orbitals of the same l and j, i.e., differing at most in their radial quantum number n, so the transformation to natural orbitals induces a change of basis on the radial functions separately within each lj space [|nlj=nan,n(lj)|nlj].

An initial NCCI calculation is carried out in the oscillator basis. This provides a scalar density matrix for the ground-state wave function, which is then diagonalized to yield the natural orbitals. The eigenvalue associated with a natural orbital represents its mean occupation in the many-body wave function. We order the natural orbitals by decreasing eigenvalue of the density matrix [21], i.e., starting with n=0 for the natural orbital with highest eigenvalue [〈𝒪0lj〉≥〈𝒪1lj〉≥...], thereby providing an n quantum number for an Nmax-type truncation scheme (see Ref. [34]).

The lowest p3/2 natural orbital obtained from the 6He ground-state one-body densities is illustrated in Fig. 2, taking an example from the NCCI calculations presented below. In a traditional shell-model description, 6He consists of two protons and two neutrons in a filled s shell (N2n+l=0), plus two neutrons in the valence p shell (N=1). In the oscillator-basis NCCI calculations, this general structure is reflected in a nearly-filled s shell. The next most heavily occupied orbital is then the neutron 0p3/2. We observe that the corresponding natural orbital [Fig. 2] receives extended contributions from high-lying oscillator shells and thus acquires a substantial large-r tail compared to the oscillator orbital, as may be expected for a weakly-bound halo nucleon.

Figure 2:
Radial wave function for the neutron 0p3/2 natural orbital (heavy curve) derived from the 6He ground state calculation in the harmonic oscillator basis, along with the contributions from individual oscillator basis functions (gray curves). The squared amplitudes P(N) of these contributions are shown in the inset. The initial oscillator basis for this calculation has Nmax=16 and ħw=20 MeV.
pic

The NCCI calculation using the natural orbital basis is no more computationally difficult than the original oscillator basis calculation. It is simply necessary, in preparation, to carry out a similarity transformation of the input Hamiltonian, as described in Refs. [35-36].

3 Results

Several experimental properties of the ground state of 6He support the interpretation that it consists of a weakly-bound two-neutron halo surrounding a tightly-bound α core [15, 16]. The two-neutron separation energy for 6He is only 0.97 MeV, out of a total binding energy of 29.27  MeV [37]. Experimentally, the onset of halo structure along the He isotopic chain is indicated by a jump in the measured charge and matter radii, from 4He to 6He. The root mean square (RMS) point-proton distribution radius rp, which may be deduced [38] from the measured charge radius, increases by ∼32% from 4He [rp=1.462(6)  fm ] to 6He [rp=1.934(9) fm] [39-41]. This increase may be understood as a consequence of halo structure, arising from the recoil of the charged α core against the halo neutrons (as well as possible contributions from swelling of the α core [41]).

The initial oscillator basis NCCI calculations from which we derive natural orbitals for 6He cover a range of oscillator basis parameters ħw=10 MeV to 40 MeV with truncations Nmax≤16, as considered in Ref. [36]. The NCCI calculations are carried out using the code MFDn [42, 43], with the JISP16 two-body internucleon interaction [12] plus Coulomb interaction.

The ground state energy eigenvalues obtained for 6He using the harmonic oscillator (dashed lines) and natural orbital (solid lines) bases are compared in Fig. 3(a). The energies from the natural orbital calculations are lower (thus, by the variational principle, closer to the true value) than those from the harmonic oscillator calculations and are also less dependent upon ħw. The improvement in convergence afforded by the natural orbitals ranges from approximately one step in Nmax in the vicinity of the variational minimum (ħw≈15–20 MeV) to several steps in Nmax towards the ends of the calculated ħw range [44].

Figure 3:
Comparison of 6He ground state properties calculated using harmonic oscillator (HO, dashed lines) and natural orbital (NO, solid lines) bases: (a) energy and (b) ground state proton radii. These are shown as functions of the oscillator basis ħw, for Nmax=10 to 16 (as labeled).
pic

For the 6He proton radius, shown in Fig. 3(b), the natural orbital basis NCCI calculations lead the harmonic oscillator basis calculations in convergence by more than one step in Nmax at ħw≈20 MeV and by several steps in Nmax at the high end of the ħw range. These radii obtained from the natural orbital calculations are also less dependent upon ħw than those obtained from the harmonic oscillator calculations.

The goal we set out to achieve is to find the true results for observables as they would be obtained in the full, infinite-dimensional space. Although full convergence is not achieved, even with the natural orbitals, the improved convergence motivates us to attempt to obtain estimates of the converged results via basis extrapolation methods [27, 29, 45, 46]. Infrared extrapolation schemes [27, 28, 47-49] are based on the premise that the solution of the many-body problem in a truncated space effectively imposes infrared (long-range) and ultraviolet (short-range) cutoffs. For bases with high enough ħw (and Nmax) ultraviolet convergence is assumed, and any remaining incomplete convergence is attributed to the failure of the basis to reproduce the long-range tail of the many-body wave function.

A basis consisting of harmonic oscillator orbitals with no more than N quanta cannot fully resolve long-range physics beyond the classical turning point L(N,ħw)=[2(N+3/2)]1/2b(ħw) [27],3 where b(ħw)=(ħc)/[(mNc2)(ħw)]1/2 is again the oscillator length, with mN the nucleon mass. The calculated energy and observables are expected to depend only on this cutoff L, approaching the true converged values as L→∞. For energy eigenvalues, it is expected that [47, 48]

E(L)=E+a0e2kL, (1)

where E, a0, and k∞ are to be deduced as fitting parameters from the results of calculations in truncated spaces. Taking →∞, we extract E as an estimate for the true energy. For mean square radii, it is expected that, letting β2kL,

r2(L)=r2[1(c0+c1β2)β3eβ], (2)

for β≫1, where r∞, c0, and c1 are similarly deduced from calculations in truncated spaces, and r∞ provides an estimate of the true RMS radius.

The extrapolated values for the 6He ground state energy and proton radius are shown in Fig. 4. We restrict ourselves to a straightforward application of (1) and (2), based on three-point extrapolation in Nmax at fixed ħw. Calculations at low ħw may not provide the assumed ultraviolet convergence, while poor infrared convergence at high ħw leads to an excessively large correction and thus poor extrapolation.

Figure 4:
Infrared basis extrapolations for the 6He ground state energy (top) and point proton radius (bottom), based on calculations in the harmonic oscillator basis (left) and natural orbital basis (right). The extrapolations (diamonds) are shown along with the underlying calculated results (plain lines) as functions of ħw at fixed Nmax (as indicated). Experimental values (circles) are shown with uncertainties. The shaded bands reflect the mean values and standard deviations of the extrapolated results, at the highest Nmax, over the ħw range considered.
pic

The extrapolated 6He ground state energies from the natural orbital NCCI calculations [Fig. 4(b)] are considerably less ħw-dependent than the extrapolated energies from the harmonic oscillator NCCI calculations [Fig. 4(a)]. The extrapolations obtained for different Nmax are also considerably more consistent (in the figure, Nmax refers to the highest Nmax in the three-point extrapolation). The extrapolated ground state energies obtained with the harmonic oscillator and natural orbital bases at ħw=20 MeV (chosen close to the variational energy minimum) and Nmax=16 are consistent with each other to within their respective variations, giving E≈-28.79 MeV and E≈-28.80 MeV, respectively.

Once the many-body calculation is under control, any remaining deviation of calculated values from nature may be attributed to deficiencies in the internucleon interaction. Comparing to the experimental binding energy of 29.27 MeV thus indicates that the JISP16 interaction underbinds 6He by ∼0.5 MeV4. (For comparison, the binding of 4He obtained with JISP16 matches experiment to within ∼0.003 MeV [29].)

The extrapolated proton radii extracted from the NCCI calculations with the natural orbital basis [Fig. 4(d)] similarly demonstrate a reduced ħw dependence and Nmax dependence, as compared to the extrapolations from the oscillator-basis calculations [Fig. 4(c)]. At the highest calculated Nmax (Nmax=16), the extrapolated rp varies only by ∼0.02 fm across the range of ħw values shown (ħw≈14 MeV to 40 MeV), and the Nmax dependence is comparable. We must emphasize that the variations in extrapolated values at best provide a rough guide to how well we can trust these extrapolated values as reflecting the true radius which would be obtained in an untruncated many-body calculation. Nonetheless, the ħw-independence and Nmax-independence of the calculations at the ∼0.02 fm level is reassuring.

Taking the extrapolated proton radius at ħw=20 MeV and Nmax=16 as representative gives rp≈1.82 fm.5 Thus, it would appear that the ab initio NCCI calculations with the JISP16 interaction, while qualitatively reproducing the increase in proton radius with the onset of halo structure in 6He, do yield a quantitative shortfall of ∼0.12 fm (or ∼6%) for the proton radius of 6He.

4 Conclusion

Describing the nuclear many-body wave function within truncated spaces is challenging due to the need to describe, simultaneously, long-range asymptotics and short-range correlations. Natural orbitals, obtained here by diagonalizing one-body density matrices from initial NCCI calculations using the harmonic oscillator basis, build in contributions from high-lying oscillator shells, thus accelerating convergence.

In the present application to the halo nucleus 6He, improvement is by about one step in Nmax near the variational minimum in ħw, and significantly more for other ħw values (Fig. 3). To put these gains in perspective, we note that an increment in Nmax results in an increase in matrix dimension of about a factor of 3–5, as seen in Fig. 1, with much larger increase in the computational costs [48]. Although full convergence is still not achieved, the calculations using natural orbitals provide improved basis parameter independence for extrapolations with respect to the infrared cutoff of the basis (Fig. 4).

The successful application of natural orbitals to ab initio nuclear NCCI calculations presented here provide a starting point for exploring ideas (some taken from electron structure theory) which may more fully realize the potential of the natural orbital approach:

(1) NCCI calculations based on natural orbitals yield improved one-body densities which can, in turn, be diagonalized to yield new natural orbitals. Natural orbitals constructed through such an iterative method can rapidly build in additional contributions from high-lying shells, thereby potentially further accelerating convergence [20, 49].

(2) An improved reference basis for the initial NCCI calculation may also boost the convergence of the subsequent natural orbital calculations. For instance, the Laguerre functions, commonly used as the starting point for natural orbitals in electron-structure calculations, also have the correct exponential asymptotics for nucleons bound by a finite-range potential.

(3) The structure of nuclear excited states can vary markedly from that of the ground state. Natural orbitals constructed by diagonalizing the density matrices from excited states, rather than from the ground state, may more effectively accelerate convergence of those excited states [20].

(4) Finally, natural orbitals are conducive to a more efficient many-body truncation scheme than the conventional oscillator Nmax scheme. The eigenvalues associated with the natural orbitals, by providing an estimate of the mean occupation of each orbital in the many-body wave function, also suggest a means of estimating the relative importance of Slater determinants involving these orbitals.

References
[1] S. C. Pieper, R. B. Wiringa, J. Carlson,

Quantum Monte Carlo calculations of excited states in A=6-8 nuclei

. Phys. Rev. C 70, 054325 (2004). doi: 10.1103/PhysRevC.70.054325
Baidu ScholarGoogle Scholar
[2] T. Neff, H. Feldmeier,

Cluster structures within Fermionic Molecular Dynamics

. Nucl. Phys. A 738, 357-361 (2004). doi: 10.1016/j.nuclphysa.2004.04.061
Baidu ScholarGoogle Scholar
[3] G. Hagen, D. J. Dean, M. Hjorth-Jensen, et al.,

Benchmark calculations for 3H, 4He, 16O, and 40Ca with ab initio coupled-cluster theory

. Phys. Rev. C 76, 044305 (2007). doi: 10.1103/PhysRevC.76.044305
Baidu ScholarGoogle Scholar
[4] S. Quaglioni, P. Navrátil,

Ab initio many-body calculations of nucleon-nucleus scattering

. Phys. Rev. C 79, 044606 (2009). doi: 10.1103/PhysRevC.79.044606
Baidu ScholarGoogle Scholar
[5] S. Bacca, N. Barnea, A. Schwenk,

Matter and charge radius of 6He in the hyperspherical-harmonics approach

. Phys. Rev. C 86, 034321 (2012). doi: 10.1103/PhysRevC.86.034321
Baidu ScholarGoogle Scholar
[6] N. Shimizu, T. Abe, Y. Tsunoda, et al.,

New-generation Monte Carlo shell model for the K computer era

. Prog. Theor. Exp. Phys. 2012, 01A205 (2012). doi: 10.1093/ptep/pts012
Baidu ScholarGoogle Scholar
[7] T. Dytrych, K. D. Launey, J. P. Draayer, et al.,

Collective modes in light nuclei from first principles

. Phys. Rev. Lett. 111, 252501 (2013). doi: 10.1103/PhysRevLett.111.252501
Baidu ScholarGoogle Scholar
[8] B. R. Barrett, P. Navrátil, J. P. Vary,

Ab initio no core shell model

. Prog. Part. Nucl. Phys. 69, 131-181 (2013). doi: 10.1016/j.ppnp.2012.10.003
Baidu ScholarGoogle Scholar
[9] S. Baroni, P. Navrátil, S. Quaglioni,

Unified ab initio approach to bound and unbound states: No-core shell model with continuum and its application to 7He

. Phys. Rev. C 87, 034326 (2013). doi: 10.1103/PhysRevC.87.034326
Baidu ScholarGoogle Scholar
[10] R. B. Wiringa, V. G. J. Stoks, R. Schiavilla,

Accurate nucleon-nucleon potential with charge-independence breaking

. Phys. Rev. C 51, 38 (1995). doi: 10.1103/PhysRevC.51.38
Baidu ScholarGoogle Scholar
[11] D. R. Entem, R. Machleidt,

Accurate charge-dependent nucleon-nucleon potential at fourth order of chiral perturbation theory

. Phys. Rev. C 68, 041001(R) (2003). doi: 10.1103/PhysRevC.68.041001
Baidu ScholarGoogle Scholar
[12] A. M. Shirokov, J. P. Vary, A. I. Mazur, et al.,

Realistic nuclear Hamiltonian: Ab exitu approach

. Phys. Lett. B 644, 33-37 (2007). doi: 10.1016/j.physletb.2006.10.066
Baidu ScholarGoogle Scholar
[13] E. Epelbaum, H.-W. Hammer, U.-G. Meißner,

Modern theory of nuclear forces

. Rev. Mod. Phys. 81, 1773 (2009). doi: 10.1103/RevModPhys.81.1773
Baidu ScholarGoogle Scholar
[14] M. Freer,

The clustered nucleus—cluster structures in stable and unstable nuclei

. Rep. Prog. Phys. 70, 2149-2210 (2007). doi: 10.1088/0034-4885/70/12/R03
Baidu ScholarGoogle Scholar
[15] B. Jonson,

Light dripline nuclei

. Phys. Rep. 389, 1-59 (2004). doi: 10.1016/j.physrep.2003.07.004
Baidu ScholarGoogle Scholar
[16] I. Tanihata, H. Savajols, R. Kanungo,

Recent experimental progress in nuclear halo structure studies

. Prog. Part. Nucl. Phys. 68, 215-313 (2013). doi: 10.1016/j.ppnp.2012.07.001
Baidu ScholarGoogle Scholar
[17] P.-O. Löwdin,

Quantum theory of many-particle systems. I. physical interpretations by means of density matrices, natural spin-orbitals, and convergence problems in the method of configurational interaction

. Phys. Rev. 97, 1474 (1955). doi: 10.1103/PhysRev.97.1474
Baidu ScholarGoogle Scholar
[18] H. Shull, P.-O. Löwdin,

Natural spin orbitals for helium

. J. Chem. Phys. 23, 1565 (1955). doi: 10.1063/1.1742383
Baidu ScholarGoogle Scholar
[19] P.-O. Löwdin, H. Shull,

Natural orbitals in the quantum theory of two-electron systems

. Phys. Rev. 101, 1730 (1956). doi: 10.1103/PhysRev.101.1730
Baidu ScholarGoogle Scholar
[20] E. R. Davidson,

Properties and uses of natural orbitals

. Rev. Mod. Phys. 44, 451 (1972). doi: 10.1103/RevModPhys.44.451
Baidu ScholarGoogle Scholar
[21] C. Mahaux, R. Sartor,

Single-particle motion in nuclei

. Adv. Nucl. Phys. 20, 1-223 (1991). doi: 10.1007/978-1-4613-9910-0_1
Baidu ScholarGoogle Scholar
[22] M. V. Stoitsov, A. N. Antonov, S. S. Dimitrova,

Natural orbital representation and short-range correlations in nuclei

. Phys. Rev. C 48, 74 (1993). doi: 10.1103/PhysRevC.48.74
Baidu ScholarGoogle Scholar
[23] T. Helgaker, P. Jørgensen, J. Olsen, Molecular Electron- Structure Theory (Wiley, Chichester, 2000).
[24] A. E. McCoy, M. A. Caprio,

Algebraic evaluation of matrix elements in the Laguerre function basis

. J. Math. Phys. 57, 021708 (2016). doi: 10.1063/1.4941327
Baidu ScholarGoogle Scholar
[25] M. Moshinsky, Y. F. Smirnov, The Harmonic Oscillator in Modern Physics (Harwood Academic Publishers, Amsterdam,1996).
[26] J. Suhonen, From Nucleons to Nucleus (Springer-Verlag, Berlin, 2007).
[27] R. J. Furnstahl, G. Hagen, T. Papenbrock,

Corrections to nuclear energies and radii in finite oscillator spaces

. Phys. Rev. C 86, 031301(R) (2012). doi: 10.1103/PhysRevC.86.031301
Baidu ScholarGoogle Scholar
[28] S. N. More, A. Ekström, R. J. Furnstahl, et al.,

Universal properties of infrared oscillator basis extrapolations

. Phys. Rev. C 87, 044326 (2013). doi: 10.1103/PhysRevC.87.044326
Baidu ScholarGoogle Scholar
[29] P. Maris, J. P. Vary, A. M. Shirokov,

Ab initio no-core full configuration calculations of light nuclei

. Phys. Rev. C 79, 014308 (2009). doi: 10.1103/PhysRevC.79.014308
Baidu ScholarGoogle Scholar
[30] P. Maris, J. P. Vary,

Ab initio nuclear structure calculations of p-shell nuclei with JISP16

. Int. J. Mod. Phys. E 22, 1330016 (2013). doi: 10.1142/S0218301313300166
Baidu ScholarGoogle Scholar
[31] M. A. Caprio, P. Maris, J. P. Vary, et al.,

Collective rotation from ab initio theory

. Int. J. Mod. Phys. E 24, 1541002 (2015). doi: 10.1142/S0218301315410025
Baidu ScholarGoogle Scholar
[32] L. Schäfer, H.A. Weidenmüller,

Self-consistency in the application of Brueckner’s method to doubly-closed-shell nuclei

. Nucl. Phys. A 174,1-25 (1971). doi: 10.1016/0375-9474(71)90999-7
Baidu ScholarGoogle Scholar
[33] M. A. Caprio, P. Maris, J. P. Vary,

Coulomb-Sturmian basis for the nuclear many-body problem

. Phys. Rev. C 86, 034312 (2012). doi: 10.1103/PhysRevC.86.034312
Baidu ScholarGoogle Scholar
[34] G. Hagen, M. Hjorth-Jensen, N. Michel,

Gamow shell model and realistic nucleon-nucleon interactions

. Phys. Rev. C 73, 064307 (2006). doi: 10.1103/PhysRevC.73.064307
Baidu ScholarGoogle Scholar
[35] M. A. Caprio, P. Maris, J. P. Vary,

Halo nuclei 6He and 8He with the Coulomb-Sturmian basis

. Phys. Rev. C 90, 034305 (2014). doi: 10.1103/PhysRevC.90.034305
Baidu ScholarGoogle Scholar
[36] D.R. Tilley, C. M. Cheves, J. L. Godwin, et al.,

Energy levels of light nuclei A=5, 6, 7

. Nucl. Phys. A 708, 3-163 (2002). doi: 10.1016/S0375-9474(02)00597-3
Baidu ScholarGoogle Scholar
[37] J. L. Friar, J. Martorell, D. W. L. Sprung,

Nuclear sizes and the isotope shift

. Phys. Rev. A 56, 4579 (1997). doi: 10.1103/PhysRevA.56.4579
Baidu ScholarGoogle Scholar
[38] L.-B. Wang, P. Mueller, K. Bailey, et al.,

Laser spectroscopic determination of the 6He nuclear charge radius

. Phys. Rev. Lett. 93, 142501 (2004). doi: 10.1103/PhysRevLett.93.142501
Baidu ScholarGoogle Scholar
[39] M. Brodeur, T. Brunner, C. Champagne, et al.,

First direct mass measurement of the two-neutron halo nucleus 6He and improved mass for the four-neutron halo 8He

. Phys. Rev. Lett. 108, 052504 (2012). doi: 10.1103/PhysRevLett.108.052504
Baidu ScholarGoogle Scholar
[40] Z.-T. Lu, P. Mueller, G. W. F. Drake, et al.,

Colloquium: Laser probing of neutron-rich nuclei in light atoms

. Rev. Mod. Phys. 85, 1383 (2013). doi: 10.1103/RevModPhys.85.1383
Baidu ScholarGoogle Scholar
[41] P. Maris, M. Sosonkina, J.P. Vary,

Scaling of ab-initio nuclear physics calculations on multicore computer architectures

. Procedia Comput. Sci. 1, 97-106 (2010). doi: 10.1016/j.procs.2010.04.012
Baidu ScholarGoogle Scholar
[42] H. M. Aktulga, C. Yang, E. G. Ng, et al.,

Improving the scalability of a symmetric iterative eigensolver for multi-core platforms

. Concurrency Computat. Pract. Exper. 26, 2631-2651 (2013). doi: 10.1002/cpe.3129
Baidu ScholarGoogle Scholar
[43] S.K. Bogner, R.J. Furnstahl, P. Maris, et al.,

Convergence in the no-core shell model with low-momentum two-nucleon interactions

. Nucl. Phys. A 801, 21-42 (2008). doi: 10.1016/j.nuclphysa.2007.12.008
Baidu ScholarGoogle Scholar
[44] S. A. Coon, M. I. Avetian, M. K. G. Kruse, et al.,

Convergence properties of ab initio calculations of light nuclei in a harmonic oscillator basis

. Phys. Rev. C 86, 054002 (2012). doi: 10.1103/PhysRevC.86.054002
Baidu ScholarGoogle Scholar
[45] R.J. Furnstahl, S. N. More, T. Papenbrock,

Systematic expansion for infrared oscillator basis extrapolations

. Phys. Rev. C 89, 044301 (2014). doi: 10.1103/PhysRevC.89.044301
Baidu ScholarGoogle Scholar
[46] R.J. Furnstahl, G. Hagen, T. Papenbrock, et al.,

Infrared extrapolations for atomic nuclei

. J. Phys. G 42, 034032 (2015). doi: 10.1088/0954-3899/42/3/034032
Baidu ScholarGoogle Scholar
[47] K. A. Wendt, C. Forssén, T. Papenbrock, et al.,

Infrared length scale and extrapolations for the no-core shell model

. Phys. Rev. C 91, 061301(R) (2015). doi: 10.1103/PhysRevC.91.061301
Baidu ScholarGoogle Scholar
[48] J. P. Vary, P. Maris, E. Ng, et al.,

Ab initio nuclear structure - the large sparse matrix eigenvalue problem

. J. Phys. Conf. Ser. 180, 012083 (2009). doi: 10.1088/1742-6596/180/1/012083
Baidu ScholarGoogle Scholar
[49] C. F. Bender, E. R. Davidson,

A natural orbital based energy calculation for helium hydride and lithium hydride

. J. Phys. Chem. 70, 2675-2685 (1966). doi: 10.1021/j100880a036
Baidu ScholarGoogle Scholar