Physics Letters B 707 (2012) Contents lists available at SciVerse ScienceDirect. Physics Letters B.

Similar documents
ds the vicinity of neutron drip l Takechi, M., Suzuki, S., Nishimur M., Ohtsubo, T., Nagashima, M., S

New Magic Number, N = 16, near the Neutron Drip-Line

arxiv: v2 [nucl-th] 8 May 2014

Nuclear Structure Study of Two-Proton Halo-Nucleus 17 Ne

Physics of neutron-rich nuclei

Measurements of the interaction cross sections for Ar and Cl isotopes

Physics Letters B 685 (2010) Contents lists available at ScienceDirect. Physics Letters B.

Interaction cross sections for light neutron-rich nuclei

Nuclear size and related topics

GSI. SINAP W. Xu, G.W. Fan (HIMAC) :,,,

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

Nuclear Physics using RadioIsotope Beams. T. Kobayashi (Tohoku Univ.)

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

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

Ozawa, A., Matsuta, K., Nagatomo,, Yamada, K., Yamaguchi, T., Momo. uzuki, T., Yoshida, K., Muranaka,., Chiba, A., Utsuno, Y.

WEAKLY BOUND NEUTRON RICH C ISOTOPES WITHIN RMF+BCS APPROACH

Reaction Cross Sections and Nucleon Density Distributions of Light Nuclei. Maya Takechi

Nuclear Science Seminar (NSS)

Lecture 4: Nuclear Energy Generation

Coulomb Breakup as a novel spectroscopic tool to probe directly the quantum numbers of valence nucleon of the exotic nuclei

Probing the shell model using nucleon knockout reactions

Role of Hexadecupole Deformation in the Shape Evolution of Neutron-rich Nd Isotopes

arxiv: v1 [nucl-th] 23 Jul 2012

Velocity-dependent transverse momentum distribution of projectilelike fragments at 95 MeV/u. Sadao MOMOTA Kochi University of Technology

Nuclear radii of unstable nuclei -neutron/proton skins and halos-

Dissociation of deuteron, 6 He and 11 Be from Coulomb dissociation reaction cross-section

Development of Secondary Electron Time Detector for Ion Beams

Direct reactions methodologies for use at fragmentation beam energies

arxiv: v2 [nucl-th] 28 Aug 2014

Exotic Nuclei. Ingo Wiedenhöver, National Nuclear Physics Summer School 7/16/2007 Tallahassee, Florida

Proton radius of 14 Be from measurement of charge changing cross sections

DIFFUSENESS OF WOODS SAXON POTENTIAL AND SUB-BARRIER FUSION

c E If photon Mass particle 8-1

Proton Elastic Scattering and Neutron Distribution of Unstable Nuclei

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

Shell Eects in Atomic Nuclei

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

S. YOKOYAMA 1;2. Abstract. Light particle-unstable nuclei were studied along the neutron. B is a possible candidate for neutron

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

arxiv:nucl-ex/ v1 21 Jun 2001

Citation PHYSICAL REVIEW C (2006), 74(5) RightCopyright 2006 American Physical So

Study of Neutron-Proton Correlation & 3N-Force in 12 C

Momentum Distribution of a Fragment and Nucleon Removal Cross Section in the Reaction of Halo Nuclei

Nuclear Spectroscopy I

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

Spectroscopy of Single-Particle States in Oxygen Isotopes via (p, 2p) Reaction

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

Improving neutron detection efficiency by using passive converters

The light nuclei spin structure from the hadronic interactions at intermediate energies

14. Structure of Nuclei

arxiv: v1 [nucl-ex] 26 Dec 2007

NUSTAR and the status of the R3B project at FAIR

Development of Ring-Imaging Cherenkov Counter for Heavy Ions

Possible existence of neutron-proton halo in

Evidence for an l = 0 ground state in 9 He

Quantum Theory of Many-Particle Systems, Phys. 540

Total Nuclear Reaction Cross Section Induced by Halo Nuclei and Stable Nuclei

Quasiparticle-Rotor Model Description of Carbon Isotopes

C NS. Direct reactions of Borromean nuclei FM50. S. Shimoura CNS, University of Tokyo

Physics of Finite and Infinite Nuclear Systems Phys. 477 (542)

arxiv: v1 [nucl-th] 8 Sep 2011

Nucleus-Nucleus Scattering Based on a Modified Glauber Theory

Measurement of activation of helium gas by 238 U beam irradiation at about 11 A MeV

Alpha decay. Introduction to Nuclear Science. Simon Fraser University Spring NUCS 342 February 21, 2011

Physics with Exotic Nuclei

The many facets of breakup reactions with exotic beams

New simple form for phenomenological nuclear potential. Abstract

Annax-I. Investigation of multi-nucleon transfer reactions in

arxiv: v1 [physics.ins-det] 9 Jan 2019

Coulomb breakup of light composite nuclei. Abstract

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

Exotic Nuclei II. Neutron-rich nuclides. Michael Thoennessen FRIB/NSCL Michigan State University

Nuclear physics: Magdalena Kowalska CERN, PH Dept.

Sub-barrier fusion enhancement due to neutron transfer

Transition quadrupole moments in γ -soft nuclei and the triaxial projected shell model

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

Neutron halos in the Na isotopes

The Proper)es of Nuclei. Nucleons

Quasi-free one-proton and one-neutron knockout reactions on 57Ni

I. 1. Nuclear Structure Study of 50 Mn by Charge-exchange (p,n) Reaction on 50 Cr

E. Fermi: Notes on Thermodynamics and Statistics (1953))

Structures and Transitions in Light Unstable Nuclei

RFSS: Lecture 2 Nuclear Properties

Central density. Consider nuclear charge density. Frois & Papanicolas, Ann. Rev. Nucl. Part. Sci. 37, 133 (1987) QMPT 540

TWO CENTER SHELL MODEL WITH WOODS-SAXON POTENTIALS

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

Observables predicted by HF theory

Alpha inelastic scattering and cluster structures in 24 Mg. Takahiro KAWABATA Department of Physics, Kyoto University

Shells Orthogonality. Wave functions

Intermediate Energy Pion- 20 Ne Elastic Scattering in the α+ 16 O Model of 20 Ne

Cross section measurements of the elastic electron - deuteron scattering

Investigation of nuclear Equation of State (EoS) is an attractive subject not only for nuclear

PHY492: Nuclear & Particle Physics. Lecture 6 Models of the Nucleus Liquid Drop, Fermi Gas, Shell

Probing surface diffuseness of nucleus-nucleus potential with quasielastic scattering at deep sub-barrier energies

The Nuclear Many-Body Problem. Lecture 2

Magnetic rotation past, present and future

First RIA Summer School on Exotic Beam Physics, August 12-17, Michael Thoennessen, NSCL/MSU. Lecture 1: Limits of Stability 1 A = 21

Minicourse on Experimental techniques at the NSCL Fragment Separators

SOME ASPECTS OF TRANSFER REACTIONS IN LIGHT AND HEAVY ION COLLISIONS

Universal Slow RI-Beam Facility at RIKEN RIBF for Laser Spectroscopy of Short-Lived Nuclei

Transcription:

Physics Letters B 707 (2012) 357 361 Contents lists available at SciVerse ScienceDirect Physics Letters B www.elsevier.com/locate/physletb Interaction cross sections for Ne isotopes towards the island of inversion and halo structures of 29 Ne and 31 Ne M. Takechi a,,t.ohtsubo b, M. Fukuda c,d.nishimura c,t.kuboki d, T. Suzuki d, T. Yamaguchi d, A. Ozawa e, T. Moriguchi e, H. Ooishi e, D. Nagae e, H. Suzuki e, S. Suzuki b,t.izumikawa f, T. Sumikama g, M. Ishihara a, H. Geissel h,n.aoi a, Rui-Jiu Chen a, De-Qing Fang i, N. Fukuda a,i.hachiuma d,n.inabe a, Y. Ishibashi e,y.ito e,d.kameda a,t.kubo a,k.kusaka a,m.lantz a,yu-gangma i,k.matsuta c, M. Mihara c, Y. Miyashita g,s.momota j, K. Namihira d, M. Nagashima b,y.ohkuma b, T. Ohnishi a, M. Ohtake a,k.ogawa e, H. Sakurai a,y.shimbara b, T. Suda a,h.takeda a, S. Takeuchi a, K. Tanaka a, R. Watanabe b,m.winkler h, Y. Yanagisawa a,y.yasuda e, K. Yoshinaga g, A. Yoshida a, K. Yoshida a a RIKEN, Nishina Center, Wako, Saitama 351-0198, Japan b Department of Physics, Niigata University, Niigata 950-2102, Japan c Department of Physics, Osaka University, Osaka 560-0043, Japan d Department of Physics, Saitama University, Saitama 338-8570, Japan e Institute of Physics, University of Tsukuba, Ibaragi 305-8571, Japan f RI Center, Niigata University, Niigata 950-2102, Japan g Department of Physics, Tokyo University of Science, Tokyo 278-8510, Japan h Gesellschaft für Schwerionenforschung GSI, 64291 Darmstadt, Germany i Shanghai Institute of Applied Physics, Chinese Academy of Sciences, Shanghai 201800, People s Republic of China j Faculty of Engineering, Kochi University of Technology, Kochi 782-8502, Japan article info abstract Article history: Received 31 October 2011 Received in revised form 14 December 2011 Accepted 14 December 2011 Available online 27 December 2011 Editor: D.F. Geesaman Keywords: Island of inversion Neutron halo Interaction cross sections (σ I ) for Ne isotopes from the stability line to the vicinity of the neutron drip line have been measured at around 240 MeV/nucleon using BigRIPS at RIBF, RIKEN. The σ I for 27 32 Ne in every case exceed the systematic mass-number dependence of σ I for stable nuclei, which can be explained by considering the nuclear deformation. In particular the σ I for 29 Ne and 31 Ne are significantly greater than those of their neighboring nuclides. These enhancements of σ I for 29 Ne and 31 Ne cannot be explained by a single-particle model calculation under the assumption that the valence neutron of 29 Ne ( 31 Ne) occupies the 0d 3/2 (0 f 7/2 ) orbital, as expected from the standard spherical shell ordering. The present data suggest an s dominant halo structure of 29 Ne and s- orp-orbital halo in 31 Ne. 2011 Elsevier B.V. All rights reserved. During the past few decades, structures of exotic nuclei have been extensively studied through a number of experiments using radioactive beams. In the 1980s, the nuclear halo structure of loosely-bound nuclei was first indicated through the measurements of interaction cross sections (σ I ) for light nuclei near the drip line [1]. This halo structure is one of the most interesting features of exotic nuclei, in which one (or two) weakly-bound valence nucleon(s) spatially extend(s) far beyond the nuclear core by means of the quantum tunneling effect. A further interesting point is that the formation of a halo structure often results from the breakdown of magic numbers and the collapse of the conventional * Corresponding author. E-mail address: m.takechi@gsi.de (M. Takechi). shell structure in exotic nuclei. A large-halo formation can occur only when the valence nucleon occupies an orbital with a low orbital angular momentum (l), because when the orbital angular momentum of a valence nucleon is relatively large (e.g. l 2) the quantum tunneling of a weakly-bound nucleon is impeded by the existence of the centrifugal barrier [2]. So far, several neutron halo nuclei have been found, and in many cases those halo structures originate from anomalous shell ordering and the valence halo nucleons in such halo nuclei unexpectedly occupy low-l orbitals [3 6]. In the 1990s, strongly-deformed nuclei were found in the heavier neutron-rich region around N = 20 [7]. This is the so-called island of inversion region [8]. For nuclei in this region the vanishing of the N = 20 magic number for neutrons was reported [9] and the inversion of amplitudes between sd-normal and pf-intruder 0370-2693/$ see front matter 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.physletb.2011.12.028

358 M. Takechi et al. / Physics Letters B 707 (2012) 357 361 Fig. 1. A schematic drawing of the experimental setup. shells has been considered along with nuclear deformation. The nature of the inversion mechanism has been extensively studied but remains unclear, and further experimental studies are needed. An advanced radioactive-beam facility enables us to explore the weakly-bound nuclei near the drip line in this island of inversion, and we can investigate the single particle orbital of valence nucleons in those nuclei through a search for the possible large low-l halo formation caused by the anomalous shell structures. The present Letter reports on the measurements of σ I for 20 32 Ne from the stability line to the vicinity of the neutron drip line. Among the neutron-rich Ne isotopes, 29 Ne and 31 Ne are particularly loosely-bound, with neutron separation energies of only 0.95 ± 0.15 and 0.29 ± 1.6 MeV [10], respectively. In the standard shell model, the 19th valence neutron in 29 Ne and the 21st neutron in 31 Ne are expected to occupy 0d 3/2 and 0 f 7/2 orbitals, but when the lowering of the intruder fp-shell is taken into consideration, the valence nucleon could occupy a 1p 3/2 -orbital state and 29 Ne and 31 Ne could have p-orbital halo structures. Moreover, the nuclear deformation could produce an elevation of the 1s 1/2 - orbital component [11,12], which could even lead to the formation of s-orbital halo structures in 29 Ne and 31 Ne. In order to clarify the mechanism of the change of shell structure in nuclei in the island of inversion, experimental evidence for the existence of halo structures in those nuclei, including 29 Ne and 31 Ne, is essential. Experiments were performed at the RI-beam factory (RIBF) operated by the RIKEN Nishina Center and the Center for Nuclear Study, University of Tokyo. An intense primary beam of 48 Ca (maximum intensity was around 100 pna) with a beam energy of 345 MeV/nucleon and Be production targets were used to produce 20 32 Ne secondary beams with energies around 260 MeV/nucleon. σ I was measured by the transmission method. In the transmission method, σ I is derived from σ I = 1 t ln( Γ Γ ), where 0 Γ is the ratio of the number of noninteracting outgoing particles to the number of incoming particles, Γ 0 is the same ratio for an empty-target measurement to correct for nuclear reactions in the detectors, and t denotes the thickness of the reaction target. The BigRIPS fragment separator [13,14] was used as a spectrometer to identify incoming and outgoing particles [15]. Fig. 1 is a schematic drawing of the experimental configuration using the second half of the BigRIPS. A carbon reaction target of 1.80 or 3.60 g/cm 2 thickness was located at the F5 dispersive focal plane of BigRIPS. Incoming particles were pre-separated and identified using the beam line between the F3 and F5 focal planes, and outgoing particles were identified between the F5 and F7 focal planes. For particle identification before and after the reaction target, magnetic rigidity, energy-loss, and time-of-flight information from ion chambers at F3 and F7 and plastic scintillation counters at F3, F5, and F7 were used. The position information from the PPACs at F3 was used to apply an appropriate emittance-cut for the incident beam so as to accurately count all of the noninteracting particles without missing them after the reaction target. In Fig. 2, σ I for Ne isotopes at 240 MeV/nucleon in the present work are plotted as a function of mass number (solid circles). Fig. 2. The present σ I data for Ne isotopes on C targets are plotted as a function of mass number (solid circles). The systematic mass-number dependence of σ I for stable nuclei (solid curve with a shaded band) and a Glauber-type calculation with the deformation effect (open triangles and dashed lines) are also shown for comparison. The open circles indicate the data for stable nuclei [16 20]. The insert shows rms radii data for stable nuclei [21] and A 1/3 dependence fitted to the data. The beam energies of the Ne isotopes differ slightly from one another depending on the separator settings. The effects on the σ I from these differences were corrected for by a Glauber calculation, which introduces a negligible increase in the error in σ I. For comparison, the systematic mass-number dependence of σ I for stable nuclei is shown as a solid curve. The dependence was calculated by a Glauber-type calculation (MOL[FM] in Ref. [20]), which gives the reaction cross section (σ R = σ I + σ inel, where σ inel is the total inelastic scattering cross section) using proper nucleon density distributions as inputs. Note that at energies above a few hundred MeV/nucleon σ I can be assumed to be nearly equal to σ R because the contribution of σ inel would be very small [37]. Inordertocalculate the mass-number dependence of σ I for stable nuclei in this mass region, the nucleon density distributions of stable nuclei are assumed to be Fermi distributions: / ( [ ]) r R(θ) ρ Fermi (r) = ρ 0 1 + exp dω, c [ R(θ) = R 0 1 + β2 Y 20 (θ) ]. Here, the quadrupole deformation parameter β 2 is taken to be zero. The diffuseness parameter c and R 0 are set so as to satisfy the condition that ρ Fermi (0) should be the typical value of the nuclear saturation density (0.17 fm 3 ), and the root-mean-square (rms) radius of the nucleon density distributions should follow the A 1/3 mass number dependence of experimental rms radii data for stable nuclei shown in the insert of Fig. 2. The rms matter radii of point-nucleon distributions for stable nuclei, shown by the solid circles in the insert, were deduced from the experimentally determined charge radii [21] by unfolding them with the proton radius 0.85 fm assuming that the proton and neutron distributions in nuclei are the same. The solid curve in the insert shows the A 1/3 dependence fitted to the data and the standard deviation of data from the solid curve is shown by the shaded area. Thus, the solid

M. Takechi et al. / Physics Letters B 707 (2012) 357 361 359 curve in the main part of Fig. 2 is the calculation using the parameters c and R 0 determined as described above and the shaded area is derived from the standard deviation of rms radii data shown in the insert. In Fig. 2, existing experimental σ I for stable nuclei (from F tomg) onactargetareshownwithopencircles[16 19] for comparison. It should be noted that the data for stable nuclei were measured at around 1 GeV/nucleon and have been corrected for the energy dependence of σ I using a Glauber-type calculation. The σ R data for 27 Al [20] are plotted without correction for the very slight difference between σ I and σ R at 240 MeV/nucleon. It can be seen that the open circles and the present data for Ne isotopes up to mass number 24 closely follow the systematics for stable nuclei. However, starting at mass number 25 the present σ I data gradually deviate from the solid curve and the deviation increases with increasing mass number. In order to study the origin of this deviation, we calculate the σ I using the finite values of the deformation parameter β 2. The β 2 = 0.4 0.6 of Ne isotopes have been deduced numerically from the intrinsic quadrupole moment obtained from the experimentally known B(E2) values or from the spectroscopic quadrupole moments [22 24,27] assuming the rotation of a uniformly charged deformed nucleus. In Fig. 2 the σ I calculated with the same diffuseness parameters c and R 0 used to obtain the systematic massnumber dependence of σ I for stable nuclei, but with each experimental β 2 for 20 24,26,28,30 Ne are shown with open triangles connected and extrapolated to 32 Ne by dashed lines. It can be seen that the trend of σ I for Ne isotopes is approximately reproduced by considering the experimentally measured quadrupole deformation of nuclei. Theory suggests that the nuclear deformation can enhance the rms matter radii which can increase the σ I in Ref. [28]. Here, the present experimental data for 20 32 Ne clarify the enhancements of σ I for neutron rich Ne isotopes beyond A 1/3 systematics and it is shown that nuclear deformations in those nuclei account for most of the enhancement. However, for the loosely-bound nuclei 29 Ne and 31 Ne, the σ I show dramatic enhancements compared to what is expected from nuclear deformation. The one-neutron separation energies (S n ) for 29 Ne and 31 Ne (0.95 ± 0.15 and 0.29 ± 1.6 MeV) [10] are as low as those for known one-neutron halo nuclei. Marked enhancements of σ I for such loosely-bound nuclei compared to those of neighboring nuclei are quite interesting, because such an enhancement often indicates the formation of halo structure, as can be seen in the data for light neutron-halo nuclei such as 11 Li and 11 Be [29, 30]. Moreover, the large difference of 86 ± 28 mb between the present σ I data for 31 Ne and 30 Ne (σ I ( 31 Ne) = 1435(22) mb, and σ I ( 30 Ne) = 1349(17) mb) is in good agreement with the recently measured one-neutron removal cross section (σ 1n )for 31 Ne on a Ctarget(79± 7mb[31]). This fact leads us to assume that 31 Ne has a structure of 30 Ne plus one loosely-bound valence neutron. The large difference also observed between the σ I for 29 Ne and 28 Ne is indicative of a similar structure. In order to examine the mechanism of the change of shell structure in nuclei in the island of inversion, we quantitatively investigated the observed large enhancements of σ I for 29 Ne and 31 Ne based on the single-particle orbital of a valence neutron in a deformed nucleus. For this purpose the present σ I data for 29 Ne and 31 Ne were compared with σ I calculated using a Glaubertype calculation with model nucleon density distributions obtained by the simple deformed single-particle-model (DSPM) calculations. The density distributions for the valence neutron of 29 Ne and 31 Ne were deduced from wave functions obtained by solving the Schrödinger equation with axially-symmetric deformed Woods Saxon potentials for given values of the one-neutron separation energy. For the 28 Ne and 30 Ne nuclear cores, Fermi-type density Fig. 3. The Nilsson diagram with colors indicating the expectation value of orbital angular momentum. The experimental β 2 for 28 Ne and 30 Ne are represented by the shaded area and the hatched area respectively. Possible orbitals which the valence nucleon of 29 Ne and 31 Ne would occupy are denoted by 1 and 2, and I and. II See text for the details. distributions were used, with parameters determined so as to fit the present σ I data under the constraint of ρ Fermi (0) = 0.17 fm 3. Fig. 3 shows the Nilsson diagram for the 28 Ne + nsystemcalculated using Woods Saxon potentials, with the diffuseness parameter c and the radius parameter R 0 taken from Ref. [32], of 0.67 fm and 1.27A 1/3 fm, respectively. The potential depth (V )is taken to be 42 MeV [32]. Colors indicate the expectation values of orbital angular momentum which show the changes in composition in each Nilsson orbital. The spherical s 1/2, d 3/2, d 5/2, g 7/2, and g 9/2 channels are included in the calculation of positive-parity orbitals, while the p 1/2, p 3/2, f 5/2, and f 7/2 channels are included for negative-parity orbitals. The β 2 for the 28 Ne core nucleus is deduced from the experimental B(E2) value as β 2 = 0.36(3) [23] and β 2 = 0.49(10) [24]. The region from β 2 = 0.33 to 0.59 is shown by the shaded area in Fig. 3. From these experimental values of β 2, possible orbitals which the 19th neutron in 29 Ne can occupy are limited to the [330 1/2] and [200 1/2] orbitals in the regions denoted by 1 and 2, respectively. (The ground-state spin-parity for 29 Ne has not yet been determined, although there are some spectroscopic data on 29 Ne [33,34].) In the insert of Fig. 4(a), the corresponding density distributions for 29 Ne are shown with the core density indicated as the shaded area. Those valence density distributions are deduced from the wave functions obtained with the experimental S n value of 0.95 MeV, where the potential depth was adjusted to reproduce the experimental S n value. The main spherical-base components calculated for the wave function 1 are 1p 3/2 (59%) and 0 f 7/2 (37%), while those for 2 are 1s 1/2 (74%) and 0d 3/2 (16%). As a reference, the spherical 0d 3/2 orbital is also plotted by the dotted curve. The angular momentum composition of the Nilsson orbital can also vary as the S n changes. Therefore, for closer comparison with the experimental data the σ I for 29 Ne on a C target are calculated as a function of S n with a Glauber-type calculation using nucleon density distributions as inputs. Here we used the MOL[FM] calculation [20,26] which gives essentially the same result as the optical-limit approximation of the Glauber theory, but also includes an approximation of the higher-order multiple-scattering effect which is important for halo nuclei (partially equivalent to the few-body effect [25]). The results assuming the Nilsson orbitals 1 and 2 and 0d 3/2 orbital for the valence neutron are compared with the present experimental value in Fig. 4(a). The experimental value is shown by a

360 M. Takechi et al. / Physics Letters B 707 (2012) 357 361 Fig. 4. (a) The σ I for 29 Ne calculated assuming the valence neutron occupies 2 (solid), 1 (dashed), and 0d 3/2 (dotted) orbitals plotted as a function of S n compared with the present experimental data shown by a solid circle with error bars. Single-particle density distributions for 1 ([330 1/2]), 2 ([200 1/2]) and spherical 0d 3/2 orbitals are also shown in the insert. (b) Same figure as (a) but for 31 Ne. Present data are compared to the σ I calculated assuming the valence neutron occupies II (solid), I (dot-dashed), and 0 f 7/2 (dotted) orbitals. The single-particle density distribution and the core density are also shown in the insert. closed circle and the shaded area represents the experimental errors for σ I and S n. It can be seen that the calculation using the 0d 3/2 density cannot reproduce the data at all. On the other hand, the present data strongly indicate that the valence neutron in 29 Ne occupies the [200 1/2] Nilsson orbital ( 2 ). It is interesting that the [200 1/2] orbital contains a dominating s-wave component which originates from the [211 1/2], namely the 1s 1/2 orbital in the spherical limit, because the contents of the orbitals are exchanged between [200 1/2] and [211 1/2] by an avoided crossing with increasing deformation, as shown by the exchange of the color in each level in Fig. 3. As a consequence, the single-particle density for the [200 1/2] orbital has an s-dominant halo tail as shown by the solid curve in the insert of Fig. 4(a). In contrast, the [330 1/2] orbital ( 1 ) contains a fairly large p-wave component with contributions from 1p 3/2 and 1p 1/2 orbitals in the continuum for the spherical shape. Though this orbital can also increase the σ I, the extent of the enhancement appears to be insufficient to reproduce the experimental data. It is probable that 29 Ne shares a similar nuclear structure with a neighboring N = 19 isotone, 31 Mg. Indeed the spin-parity of 1/2 + has been already determined for 31 Mg [35], consistent with the assumption that the Nilsson orbital [200 1/2] is also occupied by the valence neutron of 31 Mg. This suggests that both 31 Mg and 29 Ne carry an s-dominant structure for the valence neutron due to the nuclear deformation. A similar analysis has been carried out for 31 Ne. The experimental β 2 for the core nucleus 30 Ne is deduced from the experimental B(E2) data [27] as β 2 = 0.6 ± 0.2, which is shown by the hatched area in Fig. 3. Therefore, candidates for the possible Nilsson orbitals for the 21st neutron in 31 Ne are the [321 3/2] and [200 1/2] orbitals denoted by I and II in Fig. 3, respectively. (The difference between the calculated results for Nilsson orbitals of A = 30 and A = 28 is small and Fig. 3 is still helpful for considering the single particle level of the valence neutron in 31 Ne.) The main spherical-base components calculated with the DSPM using the value S n = 0.29 MeV for I are 0 f 7/2 (70%) and 1p 3/2 (22%), while those for II are 1s 1/2 (82%), 0d 3/2 (9%) and 0d 5/2 (6%). The corresponding density distributions are plotted in the insert of Fig. 4(b). As a reference, the spherical 0 f 7/2 orbital is also plotted by the dotted curve. Using the DSPM densities, the σ I for 31 Ne on a C target are calculated as a function of S n. Similar to the case of 29 Ne, a calculation using the spherical 0 f 7/2 density cannot reproduce the data at all, even considering the large experimental error in the S n for 31 Ne. Here, Nilsson orbital [200 1/2], denotedby, II contains a dominant s-wave component originating from the 1s 1/2 orbital and I contains a certain amount of p-wave component contributed by the 1p 3/2 orbital in the spherical limit. Either II or I can reproduce the present σ I data due to the large S n error. The present data support both possibilities that 31 Ne has a p-orbital halo component or a s-dominant halo structure. This result is consistent with a recent Coulomb-breakup experiment [31]. In this analysis, only the [200 1/2] and [321 3/2] orbitals are considered, given the restriction of the experimental β 2 value, but as in Ref. [36], there is another possibility for the valence neutron to occupy [330 1/2] orbital with much smaller β 2 and have a dominating p-wave halo component. This possibility is not inconsistent with the present σ I data. In either case, the present data indicate the existence of a long tail in the neutron density distribution of 31 Ne as shown by the solid curve in the insert of Fig. 4(b). In summary, σ I for Ne isotopes from the stability line to the vicinity of neutron-drip line have been very precisely measured using BigRIPS at RIBF. The σ I for neutron-rich Ne isotopes are generally and in some cases greatly enhanced compared to the mass-number dependence of σ I for stable nuclei. The general tendency of σ I to gradually increase in neutron-rich Ne isotopes is in good agreement with a Glauber-type calculation with the deformation effect taking into account the experimental deformation parameters. Besides the gradual increase of σ I for neutron-rich Ne isotopes, marked enhancements of σ I compared to their neighboring nuclei can be seen for the weakly-bound nuclei 29 Ne and 31 Ne. Such sudden increases of σ I cannot be explained by the assumption that the valence neutrons of those nuclei should occupy orbitals with d- orf -orbital angular momentum as expected from the standard spherical shell ordering. It is shown that the σ I for 29 Ne can be best reproduced by the Nilsson orbital [200 1/2] for the valence neutron which contains a dominating s-orbital component. Thus, this work is a first experimental indication of a halo structure with most probably a dominating l = 0 component in 29 Ne. In addition, the sudden increase of σ I for 31 Ne can be explained by Nilsson orbitals [321 3/2] or [200 1/2], depending on S n. This represents a low-l (= 0 or 1) orbital halo structure in 31 Ne, which is consistent with the results from a recent Coulomb breakup experiment [31]. Further experimental studies of the nuclei 29 Ne and 31 Ne, e.g. a precise mass measurement of 31 Ne, should yield a better understanding of the anomalous structure of nuclei in the island of inversion.

M. Takechi et al. / Physics Letters B 707 (2012) 357 361 361 Acknowledgements We would like to thank the accelerator staff of the RIKEN Nishina Center for providing the intense 48 Ca beam. We also wish to thank J. Miller at LBNL for his careful reading of the manuscript, and K. Hagino and K. Yoshida for making their code for singleparticle states in a deformed Woods Saxon potential available to us. This work was partially supported by the Grant-in-Aid for Young Scientists (B), 22740186, 2010, from the Japan Society for the Promotion of Science. Y.G.M. and D.Q.F. acknowledge the support by the National Natural Science Foundation of China under contract Nos. 10975074 and 10775167, and the Major State Basic Research Development Program in China under contract No. 2007CB815004. References [1] I. Tanihata, et al., Phys. Rev. Lett. 55 (1985) 2676. [2] K. Riisager, A.S. Jensen, P. Møller, Nucl. Phys. A 548 (1992) 393. [3] I. Tanihata, et al., Phys. Lett. B 287 (1992) 307. [4] J.H. Kelley, et al., Phys. Rev. Lett. 74 (1995) 30. [5] T. Suzuki, et al., Phys. Rev. Lett. 89 (2002) 012501. [6] K. Tanaka, et al., Phys. Rev. Lett. 104 (2010) 062701. [7] T. Motobayashi, et al., Phys. Lett. B 346 (1995) 9. [8] E.K. Warburton, et al., Phys. Rev. C 41 (1990) 1147. [9] P. Doomenbal, et al., Phys. Rev. Lett. 103 (2009) 032501, and references therein. [10] B. Jurado, et al., Phys. Lett. B 649 (2007) 43. [11] T. Misu, W. Nazarewicz, S. Aberg, Nucl. Phys. A 614 (1997) 44. [12] I. Hamamoto, Phys. Rev. C 69 (2004) 041306(R). [13] T. Kubo, Nucl. Inst. Meth. B 204 (2003) 97. [14] T. Ohnishi, et al., J. Phys. Soc. Jpn. 77 (2008) 083201. [15] M. Takechi, et al., Nucl. Phys. A 834 (2010) 412c. [16] A. Ozawa, et al., Nucl. Phys. A 691 (2001) 599. [17] L. Chulkov, et al., Nucl. Phys. A 603 (1996) 219. [18] T. Suzuki, et al., Phys. Rev. Lett. 75 (1995) 3241. [19] T. Suzuki, et al., Nucl. Phys. A 630 (1998) 661. [20] M. Takechi, et al., Phys. Rev. C 79 (2009) 061601(R). [21] I. Angeli, At. Data Nucl. Data Tables 87 (2004) 185. [22] J. Gibelin, et al., Phys. Rev. C 75 (2007) 057306. [23] H. Iwasaki, et al., Phys. Lett. B 620 (2005) 118. [24] B.V. Pritychenko, et al., Phys. Lett. B 461 (1999) 322. [25] J.S. Al-Khalili, J.A. Tostevin, Phys. Rev. Lett. 76 (1996) 3903. [26] B. Abu-Ibrahim, Y. Suzuki, Phys. Rev. C 62 (2000) 034608. [27] Y. Yanagisawa, et al., Phys. Lett. B 566 (2003) 84. [28] K. Minomo, et al., Phys. Rev. C 84 (2011) 034602. [29] I. Tanihata, et al., Phys. Lett. B 206 (1988) 592. [30] M. Fukuda, et al., Phys. Lett. B 268 (1991) 339. [31] T. Nakamura, et al., Phys. Rev. Lett. 103 (2009) 262501. [32] A. Bohr, B.R. Mottelson, Nuclear Structure, Vol. I, Benjamin, Reading, MA, 1969. [33] V. Tripathi, et al., Phys. Rev. Lett. 94 (2005) 162501. [34] V. Tripathi, et al., Phys. Rev. C 76 (2007) 021301(R). [35] G. Neyens, et al., Phys. Rev. Lett. 94 (2005) 022501. [36] I. Hamamoto, Phys. Rev. C 81 (2010) 021304(R). [37] A. Ozawa, Nucl. Phys. A 709 (2002) 60.