Prog. Parr. Nucl. Phys.. Vol. 35, pi. 505-573, 1995
Copyright Q 1995 Elsevier Science Ltd hinted in Great Britain. All tights resewed 0146-641oKJ5$29.00
Pergamon
01466410(95)ooo46-1
Nuclear Structure Studies from Reaction Induced by Radioactive Nuclear Beams I. TANIHATA The Institute of Physical and Chemical Research (RIKEN), 2-1 Hirosawa, Wakq Saitama 351-01, Japan
1. INTRODUCTION Since the middle of the 1980’s, high-energy beams of secondary radioactive nuclei have been used for reaction studies with short-lived nuclei. The first measurement was to determine the interaction cross sections of the reaction between unstable nuclei and stable target nuclei. Then, measurements were extended to the fragmentation cross sections and momentum distributions of Recently, new-generation facilities of fragments from radioactive projectiles. radioactive nuclear beams in GANIL (France), GSI (Germany), MSU (USA), and RIKEN (Japan) provide a wide range of radioactive beams, and various measurements have been made. Taking advantage of the high-intensity secondaries, elastic scatterings, inelastic scatterings, Coulomb excitations, and transfer reactions are now being studied. The main themes of radioactive-beam studies are the structures of nuclei far from the stability line (particularly those close to the neutron dripline) and reaction studies relevant to nucleosynthesis at various sites of the universe. In this report, I discuss structure studies of light nuclei far from the stability line after a brief overview of the facilities. Because of the fast development of this field, it is practically impossible to show all of the interests and to cover the most recent developments. Therefore, I do not try to make an exhaustive review of the studies in the field, but, instead, only mention selected studies related to the new structures; neutron halo and neutron skin.
505
506
I.
Tanihata
Studies which do not use the reactions of radioactive nuclei but only use radioactive beams as a separation method, are not mentioned in this article. Among them are the production and use of polarized nuclear beams and measurements of the electromagnetic moments. Studies using isomer beams, spectroscopic studies of g-decays, and b delayed neutron measurements are not covered. In addition, studies related to astrophysical interests are also not covered at all. Such subjects can be found in other review articles (for example, Mueller and Sherrill1993, Boyd 1993).
1.1 Production of Radioactive Beams Since the first use of a radioactive beam at the Bevalac, many facilities have been developed and radioactive nuclear beams from an energy of 1 MeV up to Also, many facilities are under 1GeV per nucleon have been supplied. construction and in the planning stages. Two different methods have been applied to the production of radioactive beams. One is a secondary beam facility in which the projectile fragmentation of high-energy heavy ions is The other is a reacceleration mainly used in combination with a separator. facility, in which radioactive nuclei produced by nuclear reactions are guided to an ion source and are accelerated again. The following sections briflypresent the existing facilities and their features. However, the principles of the production methods and details concerning the facilityare not given here.. Interested readers are asked to refer to the proceedings of the conferences (Conf.) and other material (Tanihata, I., 1989).
Secondary beam facility Projectile fragmentation in high-energy heavy-ion collisions was discovered about 20 years ago (Greiner, 1975). It was soon realized that this reaction has good characteristics for producing secondary beams of radioactive nuclei: 1. Large production cross sections of unstable nuclei, including those near to the drip line (Westfall et al., 1979; Guillemaud-Mueller et al., 1990); 2. Very small velocity broadening of the fragments (Greiner, 1975). Taking advantage of these properties, the first beam line was constructed and used in reaction studies at Berkeley (Tanihata et al., 198513). Because the velocities of fragments are almost equal, selection by the magnetic rigidity For many types of experiments the A/Z separation is gives A/Z separation. sufficient, because secondary nuclei can be identified one by one before being incident on a reaction target. High-energy beams have an advantage for this
Nuclear Sm~ctureStudies
simple separation method, not only regarding the resultant intensity of the secondary beams due to the kinematic focus, but also regarding the cleanness of the fragment momentum distribution Some types of experiments require a beam without an admixture of other nuclide in An additional addition to the objective one. separation is obtained by an energy degrader because dE/dx is proportional to 22, and which In particular, an is different from A/Z. “energy-loss achromat” method, an energy degrader technique developed at GANIL, provides a simple and efficient method of separation (Dufour et al., 1986). The new separators, the LISE at GANIL, the FRS at GSI, the A1200 separator at MSU, and the RIPS at RIKEN, use this technique. As shown in the example (FRS) in Fig. 1.1, a wedge-shaped energy degrader is placed at the dispersive focus (Geissel H. et al., 1992). The wedge degrader is so shaped that the relative momentum broadening (APIP) remains the Then, additional same after the degrader.
Production target
Dispersivefocus d degrader
I! 3
Reactiontarget
Fig. I .I Fragment separator FRS at GSI (Geissel H. etal., 1992).
-100
so
position at Final Focal Plane [mm] Fig. 1.2 Position separationsof the secondary fragments at hvo FJRSfoci. A nuclide can be sekcted by the two collimators.
separation is achieved by the last half of the separator that has the same dispersion as the first part of the Therefore, magnet sets separator. having the same dispersion can be used both with and without a degrader. This degrader technique can also be extended to other wedge shapes. Another important use is a mono-energetic degrader that is shaped so as to minimize the energy spread after the degrader. Although the beam size becomes larger with this degrader, a smaller energy spread may be essential for some experiments. The energy-
508
I. Tanihata
degrader technique is simple and effective up to several hundred MeV/nucleon. Figure 1.2 shows an example of separation at the FRS (Geissel H. et al., 1992). After collimating the beam at two foci, one nuclide can be selected. At an intermediate energy, however, since the velocity broadening is larger separation by the magnetic rigidity and the energy loss is not sufficient to select out a manageable number of nuclides. This is due to the following two facts: 1. The momentum distribution of a projectile fragment becomes relatively large and the low-momentum tail becomes both longer and larger. 2. The separation character of the an energy degrader and the magnet combination become similar to the separation character of the first-stage magnet for a lower energy (See Fig. 1.3). Recently, a new Wien filter has been installed in the LISE for further separation, as shown in Fig 1.4 (Mueller A. C. and Anne R., 1990). A clean separation of single nuclide was demonstrated for l*Li. Production target
128
130 132 134 Neutron number
136
138
Fig. 1.3 Selectivity of an energy-lossachromatseparator. The solid line shows the selection by the first half of the magnet system that select out the A/Z. The selection characteristic of the second half changes according the
b Wien filter d Smeco$daryJ
energy of a particle. From this figure, one can tell that this method has the best separation power at an energy of around5OOAMeV (Schmidt K. H. 1987).
Fig. 1.4 LISE at GANIL
The
new Wien filter has been added for better separation.
The facilities of this type are listed in Table 1-1. Among those, the RIPS at RIKEN has a unique setup (Fig. 1.5), a beam swinger in front of the production target (Kubo T. et al., 1990). The scattering angle of projectile fragments can be selected by changing the incident angle of the
509
Nuclear Saucture Studies
It then provides a tool for selecting polarized nuclei. primary beam. Polarized nuclei have been used for determining their magnetic and electric moments (Asahi, K. et al., 1992). The A1200 separator (Fig. 1.6) at NSCL is positioned immediately after the cyclotron, and delivers radioactive nuclear beams to any of the experimental system, thus giving experimenters a chance to use all of the system. Table 1-I Secondary beam facilities Facility A12OO@MSU COMBAS@JINR FR!!@GSI LISE@GANIL Notre Dame RCNP RIl?S@RIKEN
When one wishes to work on lowenergy reactions with radioactive nuclei, low-energy heavy-ion reactions have some advantages. Although it is difficult to produce beams of nuclei far from the stability line, B-unstable nuclei near the stability line can be produced with high intensity by taking advantage of the strong primary beam. With a proper selection of the incident beam
n T-m Ap/p I&. [msr] [%I Power 8 -4.3 3.0 700 - 1500 5.4 6.4 20 4360 4.5 .7 -2.5 2.0 240 - 1500 18 1.0 5.0 800 3.2 33 15 50 0.54 1.2 4 362 3.2 5.0 1 6.0 1 1500 1 5.76
Ref. Sherril(1990,1992) Artukh (1993) Geissel(1992) Mueller (1990) Becchetti (1991) Shimoda (1992) 1 Kubo (1992)
Fig. 1.5Radioactive beam-line RIPS at RIKEN.
510
I. Tauihata
and a production reaction, a beam having a small energy spread and good purity can be obtained.
A-1200 BEAM
Fig.
ANALYSIS
1.6 Radioactive beam line Al200
at the National Superconducting Cyclotron
Laboratory.MSU.
Such a facility exists at the Tandem-Van-de-Graaff laboratory in Notre Dame (Smith R. J. et al., 1990, 1991). This facility has been successfully used for studies of low energy exclusive reactions, such as elastic scatterings, transfer reactions, and Coulomb excitations. Recently, many facilities of this type are under construction. Many interesting reactions to be studied exist in nuclear A low-energy heavy-ion accelerator physics as well as in astrophysics. reserves great possibilities to open up new studies of exclusive reactions in wider isospin space. Two large projects are under consideration. One is to upgrade the NSCL facility. They are planning to combine the K500 and K1200 cyclotrons. The energies of the primary beams will be increased to 200A MeV for 160 and 9OA MeV for BU. They are expecting to increase the radioactive beam intensity by three orders of magnitude compaired to what they now have. The other plan involves the RI Beam Factory at RIKEN. There is a plan to construct a superconducting ring cyclotron with K=2000 MeV (SRC), which will be combined with the present K=540 MeV ring cyclotron. By this arrangement, the highest intensity beam is expected to be accelerated up to 500A MeV for 160 and 150A MeV for.238~. For a slightly smaller intensity, 238U can be accelerated to higher than 200A MeV. All proton dripline nuclei up to U are expected to be used with this facility. The RI beam factory also plans to construct a double-storage ring (MUSES) which will accumulate and cool radioactive beams and also store electron beam of energy up to 2.5 GeV. Studies of the electron scattering of radioactive nuclei will be possible there. These second generation facilities are expected to provide us with the opportunity to extend the horizon of nuclear physics.
NuclearStructureStudies
511
Reacceleration facility Many ISOL facilities use spallation reactions with hundred MeV. The advantage of a proton beam is of this, one usually obtains the highest production efficient acceleration of these radioactive nuclei development in many areas of accelerator physics. sources, and beam bunching.
proton energies of several its high intensity. Because yield of exotic nuclei. An is not trivial, and needs To name a few, targets, ion
Although the production of beams is more complicated, reaccelerated beams of radioactive isotopes, if produced, would in many cases be much advantageous than the recoil separator method. Best of all, they have as good a beam quality as normal accelerated beams. A good energy resolution and a good emittance are essential for many reaction studies. A drawback, however, is a limitation Because of a long extraction time from a target on the lifetime of a nucleus. and from an ion source, it is very difficult to accelerate a nucleus with lifetimes of less than one second without a significant loss of intensity. Even given this For example, reactions having restriction, many applications are expected. have been one of the most exciting importance in nuclear astrophysics applications of this method. Great efforts at Louvain-la-Neuve bear fruit (Darquennes D. et al., 1990). They used a high-intensity (up to 500 u,A) cyclotron (Cyclone30) to produce 13N through the lx(p,13N)n reaction (see Fig. 1.7). The 13N extracted from the ECR ion source is injected into the second cyclotron and accelerated. A beam of 5 x 108 /s in intensity has been used to determine the cross section of the 13N(p,$l40 reaction, one of the most important reactions for stellar nucleosynthesis. Also, a 19Ne beam of 3x108/s was accelerated at 0.65A MeV.
Productiontarget Cyclone (0.56 - 23
MeV)
(3b MeV 500 mA) Fig. 1.7 Low energy beam facility at Louvain-laNeuve. Two cyclotronsare used in tandem to produce the high quality beams of 13N and other nuclei.
Many facilities of this type have been proposed. Table l-11 list the facilities under construction. These facilities are expected to provide low-energy high-quality beams for nuclear spectroscopy and the study of reactions of astrophysical interest. In addition to the list in the table, a project has been started at CERN to construct a post accelerator after the ISOLDE.
I. Tanihata
Table I-II Reacceleration facilities under construction.
2.
RADII OF LIGHT NUCLEI
In this section the basic determination method of nuclear matter radii and the density distribution is discussed. First, the measurements of interaction cross sections and reaction cross sections are reviewed and interaction nuclear radii Then, the relation between the cross sections (and the are presented. interaction radii) and the nuclear matter density distributions is discussed. In the last sub-section, the isospin dependence of the nuclear radii is discussed. Rand
.
Cross C--tions.
.
..
The nuclear-matter radii of B-unstable nuclei were determined from an interaction cross-section measurement with high-energy radioactive nuclear beams for the first time (Tanihata et al., 1985,1985a). In this and a following series of the experiments, the interaction cross sections were measured by the The interaction cross section (al) is defined as the total transmission method. The reaction probability of one or more nucleon removal from a projectile. cross section (q& on the other hand, is defined as the difference between the total cross section (q) and the elastic cross section (c&,
Therefore, the interaction cross section is smaller than aR, and does not include target excitations without projectile excitations nor projectile excitations to particle bound states that are included in the reaction cross section. However theoretical studies based on the Glauber model (Ogawa, Yabana, and Suzuki, 1992) showed that the difference between the interaction cross section and the reaction cross section is very small at 8OOAMeV. For example, it is negligibly small for loosely bound *lLi reactions and less than 1% for well-bound 9Li reactions. However, they must be treated carefully for low-energy
513 NuclearStructureStudies experiments. Neither experimental nor theoretical studies have been made yet for energies below 1OOAMeV reactions concerning this problem. The charge-changing cross section (a~) is also often determined because of Many cross-section measurements of highsimplicity in the measurement. energy heavy ions in emulsions were based on these cross sections (Heckmann et al., 1978). It was recently measured using a radioactive nuclear beam to study the possible effect due to a difference in the proton and neutron density distributions (Blank et al., 1992). This cross section a~ is the partial cross section of the CJI. They are different by the cross sections for the removal of only neutron(s) from the projectile. Table 2-I shows a list of measurements. The measurements range from SOOAMeV down to 33A MeV. We first discuss the high-energy data, because the relation between the nuclear size and the cross section is considered to be more straightforward. Table 2-I List of the cross-section measurements for unstable nuclei 1 cross 1 E/A 1 Projectile nucleus section MeV ” 81 8-llLi or IOOO 5OS6Mn, 51958Fe, 546Q), 55,62Ni
50-o* z-5
- 12
40-o* z-2,10 m-O* Z-2- 13 790 3-8& 790 6-IILi, T-loge 790 1ILi, I I-Id&, 8-15~
01
or OR
400,800
IlLi
700 17N, 17F, &Je 40-o* z-2-
(JR
33 II&
o-2n
800 1lLi
a-2n
15
30 8He, IlLi, *4Be 80 11Li
* Beam energy integrated cross section down to 0. or : The reaction cross section determined by yray detection with 47~geometry.
514
I. Tanibata
Tanihata et al. (1985) defined the interaction radius (RI) of a nucleus by ;
(2.2) 4 where (P) indicates the projectile nucleus and (T) indicates the target nucleus. Here, the interaction radius is defined based on a simple geometrical base. In fact, it is the 35 radius of a nucleus if a black sphere is assumed and the difference between 01 and oR is neglected. It is, however, not necessary to 8 3 assume a black-sphere nucleus to define the RI by Eq. (2.2). The only necessary assumption is the ns separability of the projectile and the target radii. This separability can 2.5 be tested from data of different target and projectile combinations. It was shown (Tanihata et al., 1985a) that the separability holds well within a few % accuracy for reactions that include targets heavier than *He but lighter than Fig. 2.1 Interaction cross sections of light nudei determined by 8CKMMeV reactions. 27Al. The separability is observed to be broken for the reactions with the lightest target, such as a proton and a deuteron. However, this breakdown of the rule could be used for determining the density distribution of the neutron halo nucleus l*Li as is shown in sec. 4. 1 The separability is also expected to break for a heavy target reaction because of Coulomb interactions.
Because of the separability, the RI can be used for a comparison of the nuclear radii, even if they are determined by the different targets. However, since RI is a beam energy-dependent quantity, one should be careful when using it for other purposes. Interaction radii determined by the 8OOA MeV radioactive nuclear beams are shown in Fig. 2.1. Similar to an A*/3 dependence of the half-density radius determined by the electron scattering, the RI of stable nucleus also shows an However, the radii of unstable nuclei change drastically, Al/3 dependence. 1 A target nucleus heavier than Al is considered not to be appropriate for determining the nuclear size, because the electromagnetic dissociation cross section contributes more than a few % in the 01.
NuclearSmcture
515
Stlrdies
particulary near to the neutron drip-line. It is also seen that nuclei with same mass number show different radii. These points are discussed in the later S&iOlW.
The separability of the interaction radii have not been tested below 4OOAMeV. Therefore, a reaction model is required to connect the cross section and the nuclear radii at lower energies. 2.2
Glauber Model and Nuclear Density
The mostly used model for the interaction (a~) and the reaction cross sections (0~) is the Glauber model. When we consider the reaction in the projectile rest frame, the reaction is written as (Ogawa et al., 19921,
P(I’vo)} +T{kK&)}+ P(lO’,)}+T{I-K -n,+}.
(2.3)
The initial projectile (P) and the target (T) are in their ground states. The relative momentum is -hK. The projectile is excited by the reaction and goes to the state specified by a with a momentum transfer of fiq. The target nucleus receives a momentum transfer of -fiq and goes to state l3. It is defined that a = 0 and b = 0 stand for the respective ground states. The scattering amplitude for the reaction of Eq. (2.3) is written in Glauber theory as, (2.4) The profile function (Pii) for NN scattering depends on those transverse components of the nucleon coordinates which lie in a plane perpendicular to K. The cross section for the reaction of Eq. (2.3) is now given by,
The reaction cross section can easily be obtained by summing possible final states (a@ except for ag = 00,
oap over the
(2.6) Using the unitarity condition of NN scattering, 11- Pj’ = 1 Eq. (2.6) reduces to the familiar formula of the reaction cross section, (2.7)
516
I. Tanihata
where the phase shift function XPTfor elastic scattering is defined by,
The interaction cross section is such a probability that the projectile loses at least one nucleon after a collision with the target, and can thus be obtained by summing the ok over all possible states as except for a=O:
where the unitarity condition of NN scattering is again used. This equation relates the ol and CJR. Ogawa et al. (1992) estimated the difference between uI and a~; it was found to be less than a few percent for a beam energy higher than several hundred MeV per nucleon. A further simplification can be made if one introduces the optical-limit approximation and replaces the profile function to the NN cross sections under the zero-range limit. The 0~ can then be written as, a, = 2aJJ1-
T(~)Iw~,
(2.10)
where TO is the transmission for an impact parameter b, and is calculated from the nucleon-density distribution and the total NN cross sections as, (2.11)
where pi(s) is a z-direction integrated nucleon density distribution, (2.12)
IThe index i=P(projectile) or T(target) and okI is the nucleon-nucleon total cross sections in which indices k,Z are used to distinguish a proton and a neutron. The nucleon density distribution in the nucleus is written as ~,~(r). Equation (2.10) was used by Karol (1975) for estimating UR of heavy-ion This model was tested more collisions. He obtained reasonable success. precisely using a known density distribution (Tanihata et al., 1985a, 1991). In
517 NuclearStructureStudies the test, the proton density distributions determined by electron scattering were used. The neutron distributions were assumed to be the same as that of the protons, except for the overall normalization. It was shown that these equations (2.10, 2.11,2.12) reproduce the observed cross sections at 400A and 800A MeV within 2% for all reactions involving 7Li, 9Be, l*C, and *7Al. Therefore, this simple optical limit calculation of the Glauber model has proved However, one problem still remains. As to work well at high energies. mentioned before (Eq. 2.101, these equations are based on a zero-range In this approximation, the nucleon profile functions are set as approximation. a 6 function. The inclusion of a realistic profile function, which is determined from the nucleon-nucleon scattering cross sections, gives larger cross sections It is difficult to understand why a more realistic (Ogawa et al., 1992). calculation gives a worse fit to the data. A re-examination of the relation between the cross section and the density distribution is necessary if one wants to make a detailed comparison. 11,111 I 1 II01111 l,BOdI I 1, Coulomb The deflection has to be Y +‘2c taken into account if one is to apply the “’ _ same model to an energy lower than a z few tenths of MeV per - _ This 8 - Glauber model talc. nucleon. correction can be made classically by replacing the impact parameter (b) in Eq. IO Ian i2.10) with the nearest E (MeVln:%on) distance (b,) of the collision in the Fig. 2.2 The energy dependence of the reaction cross section and a comparisonto the Glauber model. Coulomb trajectory. Kox et al. (1987) carried out a calculation based on this model, and showed that it reproduces the observed cross section very well down to _ 10 MeV, as shown in Fig. 2.2, that shows the CJRof the **C + 1v reaction from 1OA to 2000A MeV. It is rather surprising to see that the Glauber model works so well at an energy as low as 1OA MeV, where the assumption used to derive the Glauber model is not fulfilled. This agreement could be meaningful because of a hidden reason. For example, some assumptions may not put any strong constraint in calculating the integrated quantity, such as the reaction cross section. The agreement could, however, also be accidental at such a low energy. Therefore, one should always be careful in applying the model to obtain any structure information from low energy data.
518
Tanihata the
They also parametrize contribution, as
(2.13)
where (IQ + Rsurf ) are the radii of the interaction region separated into the volume and surface contributions. They are further parametrized as, & = q,(AF3 + g3)
and
(2.14)
(2.15)
The Coulomb barrier of the colliding system (&) is given by, (2.16)
From a fit of many data of c~ from 30A to 2100A MeV reactions, the parameters It was found that rg = 1.1 fm and II =1.85 are ro, a, c were determined. independent of the projectile-target combination and the beam energy. Only c was found to depend on the beam energy, as shown in Fig. 2.3. These equations provide a simple way to compare the reaction cross sections at different energies. However, since they are purely empirical formula, one should be careful when applying them to an exotic nucleus because of a possible difference in the surface diffuseness as well as any proton-neutron density difference. When one measures CQ using a B-unstable nucleus, only rg is expected to change. Although ro includes size information of both the projectile and the targets, it provides a simple mean to study the deviation of the radius
-
10
/“-100 E/A
1000 (MeV)
Fig. 2.3 Energy dependence of the parameter c empirical reactioncrosssections.
in the
519
Nuclear StructureStudies
from r&~/3 . Figure 2.4 shows such a comparison of various data. The data at the high energy and the lower energies agree well for most of the nuclei, except for a few that are near to the neutron dripline. The large scatter of data is seen in IlLi, where data determined by a yray measurement show a much smaller ro. t-
a.
1.4
h
1
IA_81
dh
1.2 1.0
1.4?
1.0) T 14 . F IA-121 1.2 1.0
#
? 0
j
1.4;
IA-14)
APlsl
1.2
f
10 . II p
Q
cfi -1
I
0
4
4s I
1
%
Q I
I
2
3
L 1 J 312 -112 l/2
1.0 I 3t2
L S/2
Fig. 2.4 Comparison of the rg parameterfrom variaus measurementsof the beam energy from 33AMeV to 7ClOAWV.
The rg parameteris calculated from the measuredcross
sectionsusing Eq. (2.13 - 2.15).
I. Tanibata
520
Instead, the other measurement (Villari et al., 1991) at low energies gives the largest value. A recent measurement by the transmission method also gives a large cross section (Shimoura, 1991). It is, however, easily understood as follows. Since the last two neutrons are extremely loosely bound in IlLi, they are easily kicked out with a very small momentum transfer. Although this neutron removal has a large contribution on the OR, neither the target nor the fragment may be excited. Therefore, the cross section measured by y-ray emission gives a small value. However, the low-energy data still give a larger ro than that at high energy. This difference actually shows the big change in surface diffuseness, and thus, is evidence of a neutron halo in IlLi. This is discussed in Sec. 4 in more detail. The rg of neutron-rich nuclei shows some increase, as seen in the Fig; 2.4. It shows the isospin dependence of the nuclear radii, as already seen in Fig. 2.1. It is discussed later in more detail. The interaction radii RI and ro are good quantities to compare the experimental values; but are not suitable for comparing the theoretical density distributions. Although al can be accurately calculated if we know the density distribution, the density distribution can not be obtained from UI without any further assumption. Tanihata et al. (1988) used several different model distribution functions to study the density distribution, and thus the effective root-meansquare radius (R -). Three different types of distribution functions (Gaussian, Harmonic Oscillator, and Folded Yukawa) were used. The Gaussian and Harmonic Oscillator distribution have only one size parameter, so that the density distributions could be uniquely determined from the UI. It was found that the determined Rnns have same values for two functional shapes. The folded Yukawa distribution, which is used in the droplet model, has two parameters (Myers and Swiatecki, 1990). One is the size parameter (Ro) that is the radius of the sharp sphere before the folding; it becomes the half-density radius after folding of the Yukawa function. The other is the slope parameter (b) of the folding Yukawa function that is the surface diffuseness of the folded distribution. The density distribution is then written as, p(r) =
-2!I-(l+S)e4&
s sinh(s) s
,
fors
=s{T[Scosh(S)-sinh(S)]}
,
fors>S
where s = r/b and S = Ro /b with b being a decay parameter function, f@)=1e_rlh 4zb” rlb
.
of the Yukawa
(2.18)
Nuclear Structure
Studies
521
Because this distribution has two parameters, the density 0.9 distribution can not be determined uniquely from aI. o.* Instead, a locus that gives the observed oI is determined in the Ro - b plane, as shown for an example of the 9Be reaction P in Fig. 2.5-a (dashed line). o5 Also, a certain value Rrnrs ’ gives a locus in the same o 4_ plane. The solid curves in the ’ _ I 1 1 , I figure show the loci for I.5 1.0 different values of Rrnrs. As 0.9 can be seen in Fig. 2.5-a, the dashed line is almost parallel. 0.8 It therefore shows that the determined 8 R ?YllS is independently from ro and b. It was found that this situation a remains true for oI of mass number less than 12. Also, OS the Rrnrs determined by the o 4 droplet-model density * distribution is equal to the 1.0 ‘0 (fm) values obtained by the Harmonic-oscillator Therefore the Fig. 2.5 Relation between the interaction cross section and distribution. I the size and the diffuseness parameter in the droplet effective root-mean square model. The solid curves show the loci of constant rms radii of g-unstable nuclei of A radius. I 12 could be determined from the 01 rather model independently. The Rrnts values for stable nuclei (4He, 6Li, 7Li, 9Be and IT) agree well with the charge radii determined by the electron scatterings2 (Tanihata et al., 1983a). It should be noted, however, the observed similarity between the two loci does not hold for a heavier nucleus. An example is shown in Fig. 2.5-b for the 27A1 +27Al collision. As can be seen in the figure, the angle between the two loci becomes large, and thus Rrms can not be determined uniquely. Therefore ,for a heavier nucleus, we have to rely more on a reaction model to relate a cross section and a density distribution.
2The charge radii of a proton itself has to be folded in a R,, charge root-mean-squre radius.
before beeing compared with the
522
I. Tanibata I’
Bl____? +1”
11 2.5
.J.
2.2
2.5
i
IA--__ ,#$ ----_. _+_. c.---.Y
2.2
t
1-1 A-11 +
I
I
-.-/
I
-3i2 -112
I
1
2.2 1
112 312 92 Tz
2.5
*/ ,'
,pd-2 ----__ c M,’ _-
i
ii
IiI
1111’1 -I
i 0
I
Tz
-3.0 g
Fig. 2.6 Isospin dependence of the matter radii determined from the interaction cross section.
We can now see the isospin dependence of the obtained Rnns. The isospin dependence of nuclear matter radii provides a new systematics that could not be studied without radioactive nuclear beams. The observed rms radii of nucleon distributions are plotted for isobars of mass number A from 6 to 17 in Fig. 2.6. Mirror pairs of the same &spin (7Li-7Be, sLi-sB, and loBe-‘%) show equal radii within the experimental errors, suggesting that the Coulomb effect on the radii is small for these light nuclei. On the other hand, nuclei with larger isospins show larger radii, except for %e in A =9 isobars.
Nuclear Structure Studies
523
The observed isospin dependence of the radii was compared with predictions of Hartree-Fock calculations using the Skyrme potential by Sato and Okuhara (1986), as shown in the figure. Here, two kinds of potential parameter sets, SIB and SV, were employed; SIII includes a strong density-dependent interaction, while SV includes no such interaction. Both predictions gave fair agreements within 0.2 fm to the data in most cases. While the SV predicts only a weaker isospin dependence than observed, the SIII well reproduced the observed isospin dependence, except for A=9 isobars. The comparison, therefore, indicated the necessity of a strong density-dependent interaction in order to understand the observed behavior. The irregularity seen in A=9 isobars is considered to be due to a special cluster structure in 9Be, i. e., %e has a strong 2u+n configuration, and two u clusters are weakly bound by a neutron. Therefore, the density distribution is extended more than the usual nucleus. Recently, an anomalous behavior of the radius in A= 17 isobar was studied (Fig. 2.6-b) (Ozawa et al., 1994). They reported that the radius of 17Ne is larger than the T= 3/2 partner, 17N. This may suggest an anomalous structure of *7Ne. However, further study is necessary to understand the cause of the large radius. The isospin dependence of heavier nuclei up to A=25 was reported (Fig. 2.41, but only based on the form of ro (Villari et al., 1991). Generally, it is seen that the ro increases slightly for larger isospin nuclei.
3.
NEUTRON SKIN
3.1 Neutron Skins in He isotom In spite of detailed studies concerning stable nuclei having a large neutron excess (N - Z), no evidence of a thick neutron skin has been observed in stable nuclei. For example, the root-mean-square (rms) radius of the neutron distribution is larger than that of the protons by only -0.2 fm in 48Ca (N-Z = 8) and by -0.15 fm for 2081?b(N-Z = 44) if exist (Igo et al., 1979; Whitten, Jr., 1980; Krasznahorkay et al. 1991). No study, however, was carried out concerning the neutron skin of unstable nuclei until a few years ago. How a large difference can be seen between the proton and the neutron density distributions in unstable nuclei? The first empirical evidence of a thick neutron skin (-0.9 fm) was presented for 6He and sHe (Tanihata et al., 1992a). Matter radii of He isotopes derived from the interaction cross sections showed a drastic increase in the rms radius from 4He (1.57 f 0.05 fm) to 6He (2.48 f 0.03 fm) and sHe (2.52 f 0.03 fm). It was difficult, however, to draw any
I. Tanihata
524
quantitative conclusion from only these data about the difference in the density distributions of protons and neutrons. No measurement has been made for the charge or proton distribution. Concerning this problem, an important relation between the interaction cross sections and the nucleon-removal cross sections was deduced by Ogawa, Yabana, and Suzuki (1992) within the Glauber model to a loosely bound system. Suppose that 6He consists of a 4He core and two other loosely bound neutrons; and thus its ground state wave function (To) is given by,
yo = CPOQO 9
(3.1)
where @o is the ground state of 4He and the cpo is the two-neutron wave function in the ground state 6He. The final state, Ya (a# 0), includes at least two neutrons in continuum states because no particle stable excited state exists in 6He. Therefore, the final state can be expressed as, (3.2) where qkl,u denotes the wave function of the two neutrons momenta relative to the 6He, fikl and fik2.
with asymptotic
Then, the two-nucleon removal cross section (o-2.,) is obtained by summing the oaa of Eq. (2.5) with the restriction ~0, (3.3)
After some manipulation of Eq. (3.3) using relations given in sec. 2.2, we can obtain and important equation between the o-2., and the al (the interaction cross section not the reaction cross section!), a_2#He+T) - ol@He+T) - o@He+T) .
(3.4)
Here, they used the fact that since 5He is unbound, no single neutron removal This equation shows that the two-neutron removal cross channel exists. section is equal to the difference in the interaction cross sections of the “ground state” 6He and 4He. It holds when the two-neutron wave function in 6He is factorized as shown in Eq. (3.1). Although 6He was used as an example, this relation is valid for any nucleus under two conditions: one is the factorization of the wave function; the other is the lack of a single neutron removal channel. This relation, in turn, provides a way to experimentally test how well the core persists in the nucleus.
NuclearStructureStudies Table 3-I Cross sections of He iso~s
on a carbon target at 8OOA MeV
neutron
These cross sections have been measured by Tanihata et al. (1985) and 16 by Kobayashi et al. (1988). The data P(r) of sHe at 8OOA MeV with a carbon target almost satisfy this relation, as 64 shown in Table 3-1, indicating that *He 16 remains intact as a good core in 6He. When they applied it to sHe reactions, however, the relation did not hold at The 6He wave function is all. therefore strongly modified in sHe, loand thus has no identity as a core. 16 Instead, the relation
6HC
1
neutron
a_&-Ie) +o_h(*He) -ot (8He) -01 (4He>, (3.5) an extension of Eq. (3.4) under the same assumption but the *He core in sHe, holds well, as is also shown in Table 3-I. Note here that both 5He and ‘He are unbound. Therefore, sHe is well described by a *He core and four neutrons orbiting around the core.
10
16 0
1
2
R
3
4
r Vml
[fml
Fig, 3.1 Experimentally deduced nucleon density distributions of 4He and 8He. The dashed curves show the results of RMF calculations (Hirata 1991)
Under the assured assumption that *He forms a good core, the density distributions of nucleons in 6He and sHe were deduced from the ut’s for *He, 6He and sHe on C target by the optical-limit calculation of the Glauber model by Tanihata et al., (1992a). For He isotopes they assumed a harmonic-oscillator type density distribution with different size parameters for the 1s and lp orbitals. The resultant nucleon distributions of 6He and sHe are shown in Fig. 3.1. The distribution of neutrons extends far beyond that of protons, and the rms radius of neutrons (R,ms) was much larger than that of protons, Rpms;
526
I. Tauibata
(R,~s -Rpms) -0.9 fm for both 6He and sHe. It was thus shown that these nuclei have thick neutron skins. It was not due to the halo effect of shallow binding neutrons (which is discussed in the next section) because the two-neutron separation energy is 2.0 MeV for sHe. An appreciable neutron halo only appeared in a nucleus with an extremely small separation energy of the last neutron(s), i. e., 0.3 MeV for “Li, and 0.5 MeV for IBe. The 6He case was more or less marginal, however, because the separation energy is only about 1 MeV for two neutrons. 3.2
Is Neutron Skin a General Phenomenon?
Many different types of nuclear models predict a neutron skin for neutron-rich nuclei. Actually, theoreticians have been working hard to make their model not to form neutron skins in neutron-rich stable nuclei such, as 48Ca and 2o@b, because no evidence of a neutron skin has been observed in stable nuclei. As shown above in this section, very thick skins were observed in unstable neutron-rich He isotopes. What is the difference between stable nuclei and unstable nuclei in view of a neutron skin? Let us consider the related structure in the following. Hirata et al., (1991) calculated the AR”“” differences in the neutron and proton lfml 8He\ root-mean-square radii (ARrmS) using 1o _ ,,,y$$ the relativistic mean field (RMF) model. * A As can be seen in Fig. 3.2, the calculated p2w values of ARrmS for a wide mass range 0.5 k of nuclei lies in gross on a universal 48~3 straight line, and monotonically increase with AEF. AEF was defined as 0.0 - ,, A~ A* the single-particle-energy difference 160 t =Na between the last filled proton and I I I 1 Although the a& neutron orbitals. -10 0 lo AEF (Me\ correlation between AEF and ARrms is not perfect, it is natural to consider it as Fig. 3.2 Relation between the skin thickness and the Fermi energy difference between neutron beeing due to the other individual and proton. structure of the nucleus. Large ARrmS values, greater than 0.5, are seen for many nuclei with AEF larger than -10 MeV. Therefore, a neutron skin is considered to be a common structure that appears in neutron-rich &unstable nuclei. Among these, 6He, *He, and 240 are nuclei that are predicted to have the thickest neutron skins. From 0.1 to 0.3 fm of AR- are seen for stable nuclei (AEF z 0) in the figure. The present calculation predicts a slightly larger AR(0.3 fm) than the recent observation of 208Pb (-0.15 fm). (Krasznahorkay et al. 1991) They consider that a slight tune up of the parameter in the RMF model
Nuclear StructureStudies
527
would improve the result. They did not do so because their purpose was to study the global behavior of ARrm*, not to fit the individual data. A nonrelativistic mean-field calculation was reported by Lombard (1990) for light nuclei; the result shows a similar behavior.
60
SON
loo
120
15
10
5
0 r
Fig. 3.3 Neutron skin thickness and the number of neutrons in the skin for Cs isotopes by a Skyrme HartreeFock calculation.
5
10
15
[fml
Fig. 3.4
Potential and single-particle levels for stable and neutron-rich unstable nuclei by a Skyrme Hartree-Fock calculation.
The formation and growth of a neutron skin were studied in detail by Fukunishi et al. using the Hartree-Fock model with a Skyrme interaction (N. Fukunishi, T. Otsuka, and I. Tanihata, 1993). Figure 3.3 shows the thickness of the neutron skin and the number of neutrons in the skin for Cs isotopes. Thick skins including more than 5 neutrons are seen in neutron-rich isotopes of N>102. The single-particle levels and potentials for 1%~ and l*lCs are shown in Fig. 3.4. In the stable nucleus 13%~~ neutrons and protons occupy orbitals up to almost the same Fermi energy indicated by the dot-dashed lines. The proton and neutron potential are almost the same, except for the Coulomb interactions. It is this difference in the depth of the potential that allow the larger number of neutrons in stable nuclei. However, the width of the potential at the Fermi energy is almost equal for protons and neutrons. No thick neutron skin is, therefore, seen for stable nuclei, even if a large neutron excess exists. In the beta-radioactive isotope 18lCs, the difference of Fermi levels (A&) becomes large. Although the potential of a proton becomes slightly wider due to the attractive p-n interactions, a long tail of the potential gives large radii for the neutron. As can be seen in this example, a large
528
I. Tanihata
difference is generally expected between the proton and the neutron radii. A neutron skin is thus expected to be formed in neutron-rich unstable nuclei. 3
4. NUCLEON HALO IN NUCLEI 4.1
General Conceut of a Neutron Halo
In nuclei near the drip-line, the separation energy of last nucleon(s) becomes extremely small. Compared with the common 6 - 8 MeV in stable nuclei, many drip-line nuclei have a nucleon separation energy that is less than 1 MeV. The neutron-density distribution in such loosely bound nuclei shows an extremely long tail, called neutron halos. Although the density of a halo is very low, it strongly affects the reaction cross section and produces new properties in such nuclei. The basic reason of the formation of a neutron halo is simple. Let us assume a nucleus with a neutron loosely bound to an inert core. If we assume that the interaction potential between the neutron and the core is a square well, the size of the potential is that of the core nucleus itself (Fig. 4.1). The wave function of the neutron outside the potential is written as,
Dist.
Fig. 4.1 Basic concept of a neutron halo with simplest potential model.
31t should be noted here that a sharp cut distinction is difficult between the neutron skin and neutron halo. The neutron skin refers to the difference of the proton and neutron density radii. A halo “by definition” referes to the different slope factor in the density tail, which is related to the separation energy.
Appreciable neutron halos appear only in nuclei with an extremely
small separation energy of the last neutron(s), as shown in the next section; it is 0.3 MeV in case of llLi and 0.5 MeV for llBe. 2.13 MeV for 8He.
Therefore,
The two-neutron separation energies are 0.97 MeV for 6He and the terminology “skin” is more appropriate for 8He. It is
arbitrary to call the excess neutron on the surface either as a halo or a skin for 6He.
As was
shown by models, neutron skins appear in many nuclei away from the neutron drip line. They In number of neutrons can be included in a neutron skin.
also show that a considerable
contrast, a neutron halo is expected to include at most two neutrons in the last orbital.
529
Nuclear Structure Studies
(4.1) Using this wave function, the density where R is the width of the potential. distribution of the neutron is written as, p
(4 -
lyf(r)l*
(4.2)
.
The parameter K , which determines the slope of the density tail, is related to a separation energy of the neutron (E,) as,
(fiK12- 2P&,
(4.3)
where IJ is the effective mass of the system. As can be seen from these equations, when Es decreases K become smaller, and thus the tail of the distribution becomes longer. Although the surface diffuseness is known to be equal for all stable nuclei, that constancy can be understood simply as being due to the nearly constant nucleon separation energy (6-8 MeV) for stable nuclei. In general, the surface diffuseness is expected to depend on the nucleon separation energy. The neutron halo is the most pronounced case for a small separation energy (cl MeV). The momentum distribution f(pi) of the neutron is expressed by the Fourier transform of the wave function, f
W - c/(pi2+Kq
,
(4.4)
where pi is the Cartesian component of the momentum. The width of the momentum distribution is again related to the parameter K. In contrast with the density distribution, the smaller the Es, the smaller the width of the distribution. This is obviously a reflection of Heisenberg’s uncertainty principle: when the distribution in coordinate space is wide, that in momentum space is narrow. As shown above, the neutron halo is formed in a nucleus having a very small neutron separation energy. The formation of a neutron halo can be identified by a large extension of the density distribution or of an apparent large nuclear size and by a narrow momentum distribution. A sudden increase in the nuclear radii can be seen in Fig. 2.1 for l*Li, l*Be, *%e, *7B, and *7Ne. Narrow momentum distributions are observed in l*Li, l*Be, and 1% in the fragment momentum distribution and in the neutron distributions. More details concerning the momentum distribution are discussed in the following sections.
530
I. Tanihta
Except for IlBe, all known halo nuclei have a pair of neutrons in the last orbital. Therefore, these halo nuclei have a strong three body configuration that includes two loosely bound neutrons and a more bound core. Therefore, the realistic relation between the wave function and the separation energy is not as simple as a one-neutron halo nucleus. Many three-body calculations have been reported based on different types of models (Zhukov et al., 1993). However, no direct experiment has been carried out to study the three body structures of halo nuclei. Because of theoretical predictions, such as dineutron and the Effimov effect, the study of three body loosely bound system carries special interests. 4.2
Densitv Distribution of Halo Nuclei
As mentioned in sec. 2.1, the separability of the interaction radii (Eq. (2.2)) breaks in the collisions of llLi on the lightest targets (Hz and D2). However, in turn, this break indicates the sensitivity of the interaction cross sections to the details of the density distribution. It was then used to determine the density distribution of IlLi. Let us look at Eqs. (2.10 - 2.12) in more detail for the halo nucleus. We first separate the density of l*Li into the core and the halo part,
Per)-
PCb-1+ f'h
6-1.
(4.5)
After substituting Eq. (4.5) into Eq. (2.101, the interaction cross section (aI) is decomposed into a contribution from the core (QC)and that from the halo tot,), a, =2+db[1
-T,(&)J+~RJ~~~T,(&x~ -q(b)]=
a, + a,
,
(4.6)
where the relation, (4.7) was used. Although a, is exactly the interaction cross section of the core, ot, is modified byT&), the transparency of the core. Therefore, the ah corresponds to the reaction of halo neutrons by the target, but without the core reaction. By changing the target Only a large impact parameter contributes to ah. nucleus or the NN cross sections (by changing the projectile energy), a different With a smaller target or with part of the halo distribution can be studied. smaller NN cross sections, the halo density at a smaller r contributes 01. This method was applied for the high-energy (4OOA MeVand 8OOA MeV) First the interaction cross sections of IlLi with HZ, Da Be, C, and Al targets. core (9Li) density distribution was assumed to have a harmonic-oscillator distribution. Then, the halo neutron density distributions were calculated The under the assumption that the neutrons are bound in these potentials.
NuclearStructureStudies
531
potential shape is calculated from the core density distribution by folding the A least-square fit to the q‘s was made using Gaussian-shape NN interaction. the harmonic-oscillator width and the depth of the potential as the fitting parameters. Two different orbitals, Op and Is, were used for the neutron wave function, because the wave function is not known well. No two-body correlations were taken into account.
Table 4-I. Fitted parameters and root-mean-square radii of the ttLi density distribution
Halo neutrons orbital 1P
Fitting parameters a
(fm)
1.83fo.07
&, (MeV) 0.41fo.07
rms radii (fm) llLi
9Li core
halo
3.1wO.30
2.6ktO.10
4.8 f0.8
1 1.75fo.07 1.35fo.09 I3.05fo.30 2.5O~tO.10 4.8 f 0.8 1 2s * Errors indicate the Is deviation in the fits. Thus they do not include errors associated with the selection of functional shape.
Good fits were obtained for either of the orbital, but with a different depth of the potential. Figure 4.2 shows thus-obtained density distribution and Table 4 I shows the fitted parameters. The size of the core is equal to that of free 9Li, and thus shows that the 9Li core is not very much modified in IlLi. The separation energy of the neutron, which was calculated from the fitted potential, differs between the Op and 1s orbitals. The smaller value for the Op orbital indicates that it is necessary to compensate for the centrifugal barrier that does not exist for the s wave. Since no nucleon correlation, which is essential for the binding of IlLi, is considered, a direct comparison of the presently determined separation energy to the observed value (0.54fo.04 MeV) is not very meaningful. Therefore, no selection of the orbital could be made from the analysis. In fact, both orbitals give essentially the same density distribution. The root-mean-square radius of the halo neutron distribution is very large (4.8 fm). The determination of the relative contribution between the Op and 1s orbitals is still a current topic to understand the origin of the halo. From Fig. 4.2 one can visualize the halo distribution. The long tail dominates the density for r >5 fm. However the density itself is very low, l/100 of the nuclear matter density. Several theoretical model calculations based on various assumptions give a reasonable reproduction of the experimentally deduced density. They are a Green-function method (Bertsch and Esbensen. 1991), a Cluster Orbital shell model (Tosaka and Suzuki. 1990), a Faddeev calculation (Bang and Thompson 1992), a Shell model (Hayes 1991), a
532
I. Tanihata
Variational method (Zhukov et al., 1991), Hartree Fock (Bertsch, Brown, and Sagawa 1989) and others. All of them reproduce the density distribution and the binding energy within the experimental errors. Some of them are shown in Fig. 4.2-b. Therefore, we need a more accurate or variety of data to distinguish among these models (Thompson et al., 1993a).
2s wave
-
Fig. 4.2
Density distribution
density distribution
of
’ ILi.
with uncertainty.
______
The shadowed region indicates the experimentally
and Hartree-Fock
deduced
The two curves in (a) indicate the central values of the
experimental densities based on different assumptions concerning the neutron orbital. (b) present three theoretical-model
ws J-J-F
calculations (Faddeev calculation JJWHH,
The curves in
Variational
model WS,
model H-F (Thompson et al., 1993a)).
Harvey determined the density distribution of l*Li based on the combined analysis of the interaction cross section, the fragmentation cross section, and the ratio of narrow and wide components of the 9Li Pt distribution (Harvey, 1990). This analysis showed that a long exponential tail is necessary to explain the data consistently. He also mentioned that the neutron density at 8 fm from the center of *lLi, where the density is only 6 x10-4, is still very sensitive to these cross sections. It is therefore consistent with the conclusion drawn from the analysis mentioned above. Two groups used the same method to determine the density distribution of l*Be and *lLi using crI and a~ of different energies. Fukuda et al. (1991) measured the a~ of **Be+C and *lBe+Be at 33A MeV, and determined the density distribution of IlBe. They also tested the effect of a possible large
unclearSmmue
533
Studies
deformation in *lBe and concluded that only a density distribution with a long halo tail can consistently reproduce the cross sections at 33A MeV and at 790A MeV. Shimoura et al. (1991) measured the a~ with llLi at intermediate energies and fitted the data by the Glauber model for all available data. Figure 4.3 shows all CIRand q of the l1Li + C reaction. A density distribution that is essentially the same as that shown in Fig. 4.2 gives the best fit to the data including energy dependence.
I
llLi+C
with halo tail ----
without halo tail optical model Ogawa 1992
_ -----------
son --_ 2b
0
’
’
’
’
“‘I
B ‘O”
”
““I
E/A (MeV)
1000
Fig. 4.3 Energy dependenceof the 1‘Li interaction cross section. The experimental data are shownby the solid circles with error bars. None of the model fit the data perfectly.
Although a fair fit to the energy dependence is obtained, it is not as good as the fit obtained for the target dependence at high energies. Let us see a possible reason for this situation. The Glauber model has only two inputs for the calculation. One is the nucleon-density distribution, the other is the NN cross sections. Especially, only the average NN cross section, CJNN= ( app + Onn )/2 contributes to the q, because the proton and neutron distributions are the same in the target nucleus *V. In the low nuclear density (transparent) limit, the
534
I. Tauibata
energy dependence of 01 is exactly the same as that of the INN. On the other hand, no energy dependence appears in the high-density (black disk) limit. A realistic nucleon distribution should have an energy dependence somewhere in between these two extremes. Also, a very important point is the fact that two ois have equal values if the values of cram are equal (see Fig. 4.4), for example, the values at 800 MeV and 125 MeV. Thus, the (JI’Sat these two energies should have the same value. The data given in Fig. 4.3 show an asymmetry of the (~1 for high and low energies. The low energy data are consistently larger than that at high energy. Instead, the data at 8OOA MeV is lower than the line. This deviation suggests a new 1000 _ question for understanding the pp cros neutron halo. One is the reacpn cross tion mechanism and the other average is a possible new structure in neutron halo nuclei. Let us see P from the view point of the reac- 3 tion mechanism. In the i’Oo I Glauber model, Pauli blocking *g and Fermi motions are not con- V) sidered. because Fermi motion 2 broadens the effective energy 8 of the beam, it gives a negligibly small effect on oI. Kox et 10 1 I I IIIII I 1 I11111 al. (1987) studied this effect by 100 10 1000 replacing CJNN by the Pauli Energy [MeV] blocked effective NN cross section, which was determined Fig. 4.4 Nucleon-nucleon total cross sections and the averageof the pp and pn crosssections. by Giacomo et al. (1980; 1981; 1984). Pauli blocking was estimated to reduce a by -3 % at 300A MeV and -8 % at 83 A MeV. Therefore it gives the opposite effect seen in Fig. 4.3. A careful analysis of the cross sections was made by the Niigata group. They first calculated oI using Eq. (2.7) and the nucleon-nucleon profile function (Ogawa, Yabana, and Suzuki, 1992). The results are shown in Fig. 4.3 by the It shows the same behavior as the simplified optical-limit open circles. calculation shown above. Therefore, the additional assumption in the Glauber model is not the cause of the discrepancy. They then used the optical potential Their results are shown by under the Eikonal and adiabatic approximation. Although this calculation gives a good the dotted line in the figure reproduction of the cross section up to 80 MeV, no calculation was made for an energy higher than 200A MeV, because the optical potential is not known for However, it seems that their result does not reproduce the higher energy.
535
Nuclear StructureStudies
Changing cross section at higher energy if one extrapolates the line smoothly. the optical potential arbitrarily to fit the data is, of cource, not a solution to the present problem. As a conclusion, presently no model explain the energy dependence of the %i + W reaction. it should be noted that the energy dependence of *q + 1? was reproduced quite well by the Glauber model calculation [see Fig. 2.21. It may, therefore, suggest some unknown effects due to the llLi structure. One of the interesting is an excitation of the soft resonances of many different multiHowever, no other related information is yet known, and the polarities. question is still open. Size of the core in halo nuclei It is important to know how much the core of the halo nucleus is modified. Essentially, all presently available models assume that the core of a halo nucleus is not modified from its free state. As already mentioned in Sec. 3.1, the Glauber model predicts a relation between the neutron-removal cross sections and the difference in the interaction cross sections. As can be seen in Table 41, all known halo nuclei show that this relation holds reasonably well, and thus suggesting a decoupling of the core and the halo wave function in the first order. However, possible slight differences may be seen between the neutron removal cross sections and the difference in the interaction cross sections for the 6He, *lLi, and *lBe cases. They may suggest a change in the Suzuki (1992b) evaluated the component of the pure three body core. amplitude (core + n +n) from these data, and concluded that this component is 0.85 f 0.10 for l*Li and 0.87 f 0.10 for 6He. However, no study has been made for core excitation
Table 4-I. Comparison of the interaction and neutron removal cross sections* Nucleus
E/A
a-2n
0-n
aM-q(A-2)
al(A)-q(A-1)
(mb)
(mb)
(mb)
fmb)
6He
(MeW 790
189 * 14
219f8
llLi “Be
790 790
220 f 10
26Of20
*%e
790
210 + 10
169+4
*All reactions are with a carbon target.
129 + 13 182 f 71
536
4.3
I. Tanibata
Internal Momentum
Distribution and Neutron Correlation in the Halo
Studies of projectile fragmentation at high energy (1400 MeV/nucleon) have shown the following characteristics of the momentum distribution of the projectile fragments of stable nuclei: 1. The momentum distribution in the fragment frame is Gaussian. 2. This shape is the same for the beam direction (P/i) and for the direction perpendicular to the beam (I+) for the core part. 3. A spectrum is characterized by its central momentum (
) and standard deviation (o(P//)). 4 The value of
was found to be in the range of -10 to -130 MeV/c for various fragments (a negative sign for
indicates that the fragment speed is less than that of the projectile) . 4. Although the main part of the spectrum is symmetric for the P/i and Pt direction [dP//) = a(Pt) I, a longer tail is observed in the Pt direction. The width of the momentum spread (a(P//)) as well as
A,&
- AJ / (Ap - 1)
(4.8)
9
where uo=90 MeV/c (Goldhaber and Heckman, 1978). a(P//) takes its A F/2. The width (a(Pt)) of the transverse maximum value when AF = momentum distribution of the fragment is found to be equal to o(P//), consistent with an isotropic production of fragments in a frame moving at B/i= -
u1 =
t4 g> .
200 MeV/c.
Goldhaber (1974) explained the 00 based on the nucleon Fermi motion, and associated ug with the Fermi momentum (PF) as,
a$ -
(4.10)
,
4A Gaussian is exp(-P2 /20* ) and it’s three dimentional 302.
mean-square momentum
width is
537
Nuclear Structure Studies
Hiifner and where
-s
@I,
I
d3a
&
’
-dP’ = j-d2bW)jdzjd2P, W(b,z;P,,P,,- < P,, >)
,
(4.11)
where W is the Wigner transform of the one-body density matrix and contains information about the position-dependent momentum distribution of the removed nucleon. The distorting function (D(b)) contains the reaction dynamics and localizes the reaction to the nuclear surface. Under the Glauber model D(b) is essentially the same as the transmission function (T(b)) for the collision between a fragment and the target nuclei. They found that a reasonable fit was obtained only when they used a realistic nucleon momentum distribution. As a different formulation for extending the method used for stripping reactions to many nucleon removal, the momentum width of the projectile fragment is expressed by the separation energy of last nucleons, ti = 2U(&)MA, 1 -A,)
, or
(4.12)
AP
o* = 2u(&$
,
for one nucleon removal.
P Here, u is the atomic mass unit and
is an average separation energy of the removed nucleons. If we assume a two-body breakup of a projectile by separation energy (E), Eq. (4.4) gives the momentum of the fragment under the sudden approximation. By approximating the Lorentian by the Gaussian that has the same second derivative at the peak, one can obtain Eq. (4.13). The only difference is replacing by E. It therefore gives the same conclusion as shown in Sec. 4.1, except for a tail part of the distribution. The tail of the momentum distribution is strongly affected by D(b). One can obtain Lorentian tail only when D(b)=1 for all value of the impact parameter b. If D(b)<1 the tail of the Lorentian becomes shorter. The P/l distribution of fragments from halo nuclei l*Li, “Be, and 14Be were recently measured at NSGL and at GSI (Orr et al., 1992; 1993; Kelley et al., 1993; Riisager 1993). All of them show a width o(P//) - 20 MeV/c, much narrower than that observed in the fragmentation of stable nuclei, as shown in Fig. 4.5 as an example.
538
I. Tanihata F’l”“I”‘J
I ,
100
150: (a)
I 1 I I , I I I I -
: (c) o = 20.9 f 0.8 MeVlc -
80 -
o - 21.2 f 0.7 MeVlc _
p
0 -
18.7 * 0.8 MeVlc :
o - 20.9 f 0.6 MeVlc
(MeVk)
Fig. 4.5 Parallel momentumdistributionof 9Li fragmentsfrom 11Li reactions.(a) Be (b) Nb, and (c) Ta targets. (c) High-acceptance mode using a 9Be target. Single Gaussian fits were made for (a), (b), and (c). A tw@Gaussianfit is shown for (d). The wide Gaussianhas u-1 I2fl3
MeVk)
As discussed above, the a(&) measure the same if a high-energy projectile is used. The transverse-momentum distributions of fragments from 1% and llLi reactions with a carbon target were measured at 790A MeV (Kobayashi et al., 1988, Tanihata et al., 1992). Figure 4.6 presents the transverse-momentum distribution of (a) a loBe fragment from *IBe + C reaction and (b) a 9Li fragment from *lLi + C reaction both at 800 MeV/nucleon. Both of the data show a very narrow peak on top of another wider peak. The fitting of the momentum distribution by two Gaussians gives a width onarrow= f 4 MeV/c and In 9Li, onarrow= 21 f 3 MeV/c awide= f 7 MeV/c for the *oBe spectrum. and awide= 80 f 4 MeV/c. Although the separation of the momentum distribution into two Gaussians is somewhat arbitrary, it was shown that the fit neither by a Gaussian nor by a Lorentian for the entire range of momentum is In conclusion, the poor. We will come back to this discussion later again.
539
NuclearSpucture Studies small values of a(P//) and a(Pt) indicate that the removed small momentum fluctuation, and thus has long density tail.
neutron(s)
has a
The narrow width of loBe is 600 l”“l”“l”“~““I’c : 1lBe + C --> loBe + X consistent with a value of 21 MeV/c that was estimated from the separation energy of 400 a neutron (Es (n) = 503 keV) using Eq. (4.12). It is more % complicated for IlLi, because u two neutrons are in the same 200 loosely bound orbitals. However, if two-neutron cluster removal is assumed, 0 the observed narrow width 400 0 200 -200 -400 gives = 0.34 f 0.16 MeV, I”“I”“I”“. which is consistent with the -11Li+C-->9Li+Xat 8OOAMeV separation two-neutron energy: AE[eLi + 2n) - (IlLi)] 1000 = 0.34 f 0.10 MeV. Therefore, both spectra are consistent with the expected narrow 0 - 2 If 3 MeVlc momentum distribution of weakly bound last neutrons, 500 and thus show the formation of a neutron halo.
a
The momentum distribution in of neutrons the fragmentation region may 0 also carry halo information. 400 0 200 -200 -400 Moreover, the correlation of Transversemomentum (MeVk) neutrons in the halo may be studied by the comparison of Fig. 4.6 Transverse momentum distribution of the fragments from halo nucki. the neutron spectrum and the related fragment spectrum (Tanihata et al., 1992). Recently, many measurements of neutrons from 6He, 8He, 9Li, **Li, 1113e,and We have been reported (Kobayashi 1993). However, as already mentioned above, the nucleon distribution does not behave as that of a fragment, because of a complication in the reaction mechanism, such as evaporation p’ocesses. It has already been pointed out by Hiifner and his collaborator that one has to study the fragment momentum distribution, not the nucleon distribution, to
540
I. Tanihata
obtain the internal motion. Neutron Pt distributions from various projectiles at 800A MeV are shown in Fig. 4.7. Let see first the 9Li + C -> n + 8Li + X reaction (neutrons are detected in coincidence with aLi), the Pt distribution is wide and has a width ot = 110 MeV/c. It is consistent with the distribution of neutrons of diffractive scattering on a C target (n+C) with an initial momentum spread on the order of the Fermi motion of a neutron in the projectile. The curve in this figure shows thus calculated distribution.
1041’
i..‘.1‘..‘1““1”“1’
1”“I”“l”“l”“l‘
8He+C -> n+6He+X
: 6He+C -> n+4He+X
&-
i
:
.
p\
‘\:
. l lLi+C -> n$Li+X _i>,
..= /
7 i
‘. . a
:
Taransverse Mometum P, [MeV] Fig. 4.7 Transverse-momentum distributions of neutrons from fragmentation of neutron-richnuclei at 790A MeV (Kobayashi 1993).
This component is observed in all of the data shown in Fig. 4.7, indicating that it is a general process considered as above. In addition, a narrow peak is seen in each of the spectra on top of this component for others. In these reactions, a projectile has two valence neutrons. One of them become unbound if the other is removed, and is thus emitted simultaneously without any further collisions. This is why one could expect that this narrow spectrum can give information of the internal momentum of the halo neutron.’ Figure 4.8 shows the neutron separation energy (S, or S& dependence of the width. The a0 of the fragments in one- and two-neutron removal shows a monotonic rise for larger Sxn. The
541
Nuclear StructureStudies
solid line in Fig. 4.8-a shows the calculated value by Eq. (4.13), assuming a twoneutron cluster for two-neutron removal (--s&. The line reproduces the data well. On the other hand, the width (o) of the neutron distribution does not follow the same tendency. The width for sHe and 9Li is as small as that for IlLi, which thus show that the other reaction mechanism is important. I I 1 1
I 1 1 1
A sequential decay process is G I a Fragments considered to be important for $ d) 100 the neutron spectrum. In this q process, one neutron is removed by the interaction of a neutron with the target (see Fig. 4.9). The recoil momentum distribution of the “A-l” system is given by the one-neutron separation energy S,. Then a particle unbound “Al” decays by emitting a second neutron having a recoil momentum given by the decay Q value. The mean-square momentum I b Neutron width of the second neutron is 80given by,
a*(n) =
2us, A(A - 1)
,I,,
II
direct breakup
+zuQ(A-2) --. 3
A-l (4.14)
The first term is usually small be0 _ ----__ cause of the factor l/A (A-l). ; two-step breakup Therefore, the width of the neutron spectrum is mainly deterOt,, mined by the decay Q value. Neutron Separation Energy MeV} Using the known excited states shown in Fig. 4.9, Kobayashi ob- Fig. 4.8 Width of the momentum distributions of tained the widths indicated by fragments and neutrons from the reaction of neutron rich nuclei. Solid lines present the prediction by the the dashed line in Fig. 4.8-b. single step direct reaction and dashedlines are for two Although some discrepancies stepprocesses. remain, the over all tendency of -For a more realistic fit one has to include other the o is reproduced well. channels, such as the direct break up and sequential decay from higher excited states. ‘The contributions from these processes give a larger value of o, and thus are consistent with the deviation of the data from the dashed line. However, no realistic calculations including all of those effects have yet been made.
542
I. Tanibata
In contrast to the neutron spectrum, the o for fragments is rather insensitive to this sequential decay process. It is because o is expressed as, (4.15) In this case the first term dominates in most of the situations. In conclusion, it was found that the two-step process (fragmentation and evaporation) is important. In this process, although the fragment momentum spectrum is determined mainly by the fragmentation process, the neutron spectrum is determined mainly by the evaporation Q value.
Scatteredneutron 10 8 6
MeV
:-,-
7Li+2n
Fig. 4.9 Two-step process of fragmentation reactions and the related levels for the emission of neutrons.
The neutron Pt spectra were also measured at lower energies, as shown in Fig. 4.10. They essentially show the same behavior. However, a difference can seen in the spectra with heavy (high-Z) targets. The neutron distribution from the llBe+C ->n+lOBe+X reaction shows no narrow peak, because llBe is one
543
Nuclear StructureStudies
neutron halo and the neutron is scattered by the reaction. However, a narrow This indicates the peak rapidly develops when the target becomes heavier. progressive importance of the Coulomb interactions, because the nuclear It also indicates that the probability of reaction does not change strongly. exciting the projectile by a strong interaction is small. Therefore, the following three different mechanisms are involved in the reaction: 1. scattering or diffraction of halo neutrons, 2. scattering of the core loBe, and 3. Coulomb dissociation. Until now, no firm determination of the relative importance of these mechanisms have been made quantitatively. The Coulomb interaction is discussed in Sec. 5.
4.4 Correlation between neutrons in a halo.
two
41 MeV/u
““0
Be target 0 0
I’ ‘,
4 \
“Be
o
(%a.
.
(“&.‘&+n)
n)
41 MeV/u
,, ‘*
“Be
Au target
Except for 11Be, two neutrons form halos in nuclei. In all cases, the nucleus with one less neutron is not particle stable. It is therefore obvious that the correlation between two halo neutrons is essential to make these nuclei bound. The paring energy, for example, was studied for .I inter-media &mass nuclei. It was Neutron Angle (deg.) found that the paring energy decreases as the neutron number Fig. 4.10 Transverse momentum distribution of neutrons increases. If one extrapolates this from 11Be reactions. result to the neutron dripline, the expected paring energy would be positive, and thus unrealistic. On the other hand, the odd-even regularity of the particle stability observed near the dripline indicates the importance of the correlation. Is it due to the same mechanism observed as a paring correlation near the stability line? Or is it a different correlation, such as a di-neutron cluster?
544
I. Tanihata
An extraction of the strength of the correlation was first attempted by making a comparison of the one-neutron and two-neutron momentum distributions. The sum momentum of two neutrons (PsUm) is expressed in terms of the momentum of the individual neutron momentum, P sum - Pt + P2 .
(4.16)
Then, the broadening of these momenta is obtained by averaging the square of the momentum. Thus,
- +
+2 .
(4.17)
Here, the correlation appears in the last term, . Since two halo neutrons are identical, and cP& are equal. The value of
12C+6He-> 4He+X -
2.0
(p&2 shell model
-
I
di-neutron cluster model experiment r /
1.5 1.0
\
I’ \ p *\’ I *+
-
0.5 0 -200
-100
0
loo
200
Fig. 4.11 Transverse momentum distribution of a 4He fragment from a 6He reaction. Neither of the simple models reproduce the data if the observed radius of 6He is fitted.
54.5
Nuclear Structure Studies
making a hybrid model, which is a linear combination of the cluster orbital shell model and the di-neutron model, they could fit the experimental data quite well. Although this fitting was rather artificial, it shows the importance of selecting the correlation of neutrons. In fact the density distribution of halo neutrons is related not only to the singleThe particle wave function, but also to the position correlation . momentum distribution is then related to , as already mentioned. Because these two quantities are inter-related by wave mechanics, one can expect to extract the correlation of halo neutrons in a rather model-independent way from a comparison of the density distribution and the fragment However, no serious analysis has yet been made. momentum distribution. For 6He, the microscopic three-body calculation is expected to provide accurate information, because the core 0.08 4He is extremely tightly bound. Such a calculation has been reported by many authors; they all reproduce the momentum distribution of the 4He fragment well. Here, as an example, the results of three-body calculations by Zhukov et al.. [Zhu931 are presented. Figure 4.12 shows a twoparticle correlation density plot of The two-neutron density 6He. shows two separate peaks: one corresponding to a cigar geometry and the other corresponding to a di-neutron configuration. Although no direct information of such a correlation has been observed yet, it would be interesting to see experimentally whether such configurations are mixed in a nucleus. In the same figure the calculated Pt distribution of 4He from a fragmentation of 6He is shown (solid -250 curve). It shows a quite good fit to the data. Recently measurements of all three final state particles (9Li and two neutrons) after the fragmentation of **Li were made at MSU and at
6He Spatial correlations
10
--_ 6He
-150 4He
-> 4He
breakup at 400A MeV
-50
50
150
_I 250
transverse momentum
Fig. 4.12 Two-neutron distribution of the 6He wave function
in a three-body
momentum model.
distribution
calculation of 4He
and the
in the same
546
I. Tanihata
RIKEN and two-neutron relative-momentum spectrum was studied. However, the spectrum shows a statistical behavior presenting a importance of the final-state interactions. The difficulty here is due to the fact that the interesting relative momentum of two neutrons is much smaller than the inverse of the range of the neutron-neutron interactions. Therefore, an experiment that has weaker final-state interactions, such as (p, pn) quasi-free nucleon-nucleon scatterings, would be suitable for studying the correlation of the neutrons in a halo state. This is because the kicked out neutron has a momentum that is very far from the momenta of both the fragment 9Li and the partner neutron. However, no such measurement has yet been reported. In spite of several trials, no definite conclusions between halo neutrons have been obtained so far.
concerning
the correlation
In addition to the data shown above, many measurements have been reported for **Li reactions. These are summarized in Table 411 for the two-neutron removal cross sections and in Table 4111 for the momentum distributions (they were compiled by Riisager 1993).
Table 4-11 Two-neutron removal cross sections for 1lLi in barn.
1. Ieki, K. et al., (1993), 2. Riisager, K. et al., (1992). 3. Blank, B. et al., (I 993). 4. Kobayashi, T. et al.. (1989).
547
Nuclear Struuure Studies
Table 4-U Deduced width (FWHM) of the momentum distribution for ltLi reactions.
1. Ieki, K. et al., (1993). 2. Riisager, K. et al., (1992). 3. Humbertet al., to be published (1994) 4. Kobayashi et al., (1993). 5. Orr, N. et al., (1992). 6. Oeissel, H. et al., (1993). 7. Kobayasbi, T. et al., (1988). 8. Tanihata, T. et al.. (1992).
.
Other related cross sectrons (p
.
.
cross sectrons.I
Does the proton distribution change when a neutron halo is developed? Since all of the measurements for unstable nuclei so far use strong interactions, no direct information on the charge distribution or proton distribution is available. The neutron halo is considered to be formed only by neutrons from several indirect measurements. The separation energy of a proton in a neutron-rich nucleus is large so that the Also, as wave function is expected not to have a long asymptotic tail. discussed in the previous section, the narrow distribution of fragments was one of the decisive evidence of the halo. This narrow width is observed only in fragments in which only one or two neutrons are removed, depending on the single-neutron halo or two-neutron halo, respectively. Other fragments, in which at least one proton is removed, do not show such a narrow distribution. Therefore, a proton does not show a halo-like distribution in neutron halo nuclei.
548
I. Tanihata
Recently, Blank et al. (1992) measured the charge-changing cross I section of Li isotopes. If we believe the geometrical model, the charge changing cross section gives the lOOOoverlapping size of a proton. In 800practice, however, a proton may also be emitted from an excited fragment that is produced by neutron removal. This process gives 400an additional cross section to the charge changing. Therefore, the 1 I I charge-changing cross section is 200 8Li 9Li 1lLi considered to give the upper limit of the proton radius. The measured Fig. 4.13 Charge changing cross section and the charge-changing cross sections do reaction cross sections of Li isotopes at 80A MeV on C target. (Blank et al., 1992) not change from 7Li to IlLi, as shown in Fig. 4.13. It therefore indicates that the proton distribution does not change very much from 8Li to 11 Li.
bi
Table 4-11 Electromagnetic moments of Li isotopes and 8B. nucleus
Spin
Magnetic Quadrupole moment (n.m.) moment (mb) 7Li 3/2 3.2564 -40.0 f 0.3 8Li 2 1.6536 31.1 f 0.5 9Li 3/2 3.4391 f 0.0006 -27.4 f 1.0 1lLi 3/2 3.6678 f 0.0025 -31.2 f 4.5 8B 2 1.0355 68.3 f 2.1 (Arnold et al., 1992, Minamisono et al., 1992)
The magnetic moment and the quadruple moment of a nucleus provide a good In particular, the to study the static-electromagnetic properties of nuclei. quadrupole moment is proportional to
Nuclear Strum
Studies
549
All of the measurements so far are consistent with the persistence of the proton distribution before and after the formation of a neutron halo. Therefore, the known strong attractive p-n interaction does not affect the proton distribution to any visible degree. However, it is stlll an open question as to whether the proton distribution is completely decoupled from the halo or not.
4.6
Proton Halo?
At the limit of the bound-proton rich side, many nuclei have separation energies of less than 1 MeV. Do they have proton halos? Recently, the large quadrupole moment of sB was claimed as an evidence for a proton halo (Minamisono et al., 1992). They compared the newly measured quadrupole moment of 88 and that of 8Li. Under the assumption that the effective charges of the proton and neutron are equal to those of neighboring nuclei, they concluded that the rms radius of the proton distribution should be large,
550
I. Tauibata
not extend as much as that of neutrons. In fact the tail extends only up to the one that correspond to about Es=1 MeV of the neutron distribution. The Therefore, the density Coulomb barrier in this case is about 1 MeV. distribution is same as that with an effective separation energy, E,eff = lMeV, that is the depth counted from the barrier height.
r(fd
r (fm)
Fig. 4.14 Single-particle density distributions of a nucleon in a WoodsSaxon potential.
NuclearSmtcture Studies
551
if we compare the neutron density tail for different 1, the tail is shorter for higher 1 orbitals. It is also true that the behavior of the tail is similar for the same effective separation energy. The addition of a Coulomb barrier to a high I orbital is again shown in the figure.
Now,
2.2 Now let us relate the above mentioned tendency to the 2.0 observed neutron halo nuclei. First of all, there is 1.8 only one example of a single neutron halo, **Be. It shows 1.6 a pronounced halo, as already discussed in sec. 4. It 1.4 is expected because the last neutron has lq 12 as the main component of the wave 1.2 function On the other hand, I?, a candidate of a neutron 1.0 112 712 halo, shows no enhancement -312 -l/2 312 512 Ne F 0 N C B of ol in a measurement at *z intermediate energies (Fig. 4.15). As already shown in Fig. 4.15 The radius parameter rg determined from various sec. 4, the enhancement of q measurements. (Ozawa et al., 1994) No big enhancement is appears sensitively at a suggested in spite of its small neutron separation energy. The low energy data of t7B suggests a large enhancement of lower energy. It therefore the cross section. indicates that *7C may not have a large neutron halo in spite of its small separation energy (0.729 MeV). The single-particle orbital of the last neutron is expected to be 065/z in *7C, and is thus consistent with the large centrifugal barrier. All other known cases (6He, l*Li, IdBe), and the possible case of 17B have two neutrons coupled to jn = 0. In a two neutron halo, although single-particle orbitals include I= 0, 1, and 2, the centrifugal barrier sensitively depends on the configuration of these two neutrons. For example, if they are coupled to a OS di-neutron, no centrifugal barrier takes place. On the other hand, if they are in a pure shell-model orbital without any correlation, the centrifugal potential is the same as that felt by a single particle in the same orbital. It, therefore, requires a detailed microscopic calculation to determine the real strength of the barrier. However, the existence of a halo suggests that the centrifugal barriers for these nuclei are smaller than expected from the single-particle orbital alone. The situation is different for proton halo candidates. The candidates, sB, 12N, and 17F, have a single proton in the last loosely bound orbitals. Therefore, both the Coulomb barrier and the centrifugal barrier come into play, and thus the
552
I.Taniha
tail of the density distribution does not extend very much. It is therefore consistent with a small or no enhancement of the interaction cross section (q). The 17N nucleus is an exception in which two protons exist in the last orbital. In fact, an enhancement of aI has recently been observed in 17Ne + C collision (Ozawa et al., 1994). Thus may indicate that the centrifugal barrier is very low These situations are in this nucleus, and a proton halo may be developed. summarized in Table 4111. Table 4-111Halo candidates and their characteristics
candidates
S, or S, (MeV)
Sn or S2P
01
Orbital
enhancement
(MeV) 6He 1lLi V3e 14Be 17B 19B 17C W 8B 9C 12N 17F *7Ne
0.97
1.051 0.503
0.247 1.28 (2.45) (0.87)
0.739 (0.22) 0.14 1.299 0.601 0.600 0.96
@p3/2)2 (Opl/2orh/2)2 h/2 (Op1/20rh/212 G&/2)2 @k/2)2 f-W2 O&i/z
Yes Yes Yes Yes prob. yes
oP3/2
no
no
@p3/2)2
OP1/2 Od5/z 1so
(O&nor
ls1/2)2
prob. no no yes
All known halo candidates behave as expected based on the barrier effects. Those candidates with no barrier l*Be clearly shows the halo. In addition, all neutron halo candidates with two neutrons in the last orbitals show an enhancement of aI. Among them is the well-established halo nucleus IlLi. All candidates that have a single particle in the last orbital with 1>1 show no enhancement of a~. However, the barrier height is sensitively determined self consistently by the wave function. Therefore, the barrier height depends on the density distribution at the surface region of the nucleus, and becomes lower when the wave function has a longer tail.
Nuclear Struchne Studies
553
5. Soft Mode of Collective Excitation Because of the excess of neutrons on a surface, either the skin or the halo, new types of excitation modes can be expected. In particular, when a core and surface neutron are decoupled as
(5-l) some of the collective strength should be shared with the relative movement of the core and the surface neutrons. Although several predictions have been made for higher multipole excitations, such as E2, the study of such collective excitation is in its very primitive stage, and experimental searches are made only on the El mode of excitation.
10
A 100
Fig. 5.1 Target mass dependence of the interaction cross section and the two-neutron removal cross sections of neutron-rich nuclei. Large enhancements are seen in heavy-target reactions.
Electromagnetic dissociation(EMD) provides important information on collective excitation of nuclei. For stable nuclear projectiles, observations of the EMD process at energies from 1 (Heckman and Lindstrom, 1976; Olson et al.,
554
I. Tanihata
1981; Mercier et al., 1986; Hill et al., 1988) to 200 GeV/nucleon (Hill et al., 1988a; Price et al., 1988) have been reported, and enhancements of fragmentation due to the EMD process have been clearly seen, particularly in the single-nucleonremoval channels on high-i! targets. This phenomenon was interpreted as resulting from the excitation and decay of an El giant dipole resonance (GDR) in a projectile by virtual photons associated with a rapidly changing Coulomb field. Kobayashi et al. (1989, 1989a) observed the EMD process from the target dependence of the interaction cross sections (or) for the first time for unstable nuclei. Figure 5.1 shows the target mass dependence of the interaction cross sections, as well as the neutron (single- or two-) removal cross sections, for selected neutron-rich nuclei. The estimated nuclear parts of the cross sections are also shown in the figure by solid curves for a1 and dashed lines for
I
- . . -*“‘I
’ ’ ”
3 2 _ ,Iti 3 8He zro i f ‘g “Bel I”Be f 12Be v, ! 101 ’ PLi 8 . ,a_B,MD/z; * ,oO l e:m/zzp ‘% !! f I..: I * . . . . . . . I i IO 1 Separationenergy (MeV)
In order to compare the observed EMD Fig. 5.2 Separation energy dependence of the EMD cross sections. The EMD cross cross section of neutron-rich nuclei with a section increases rapidly with a decrease typical cross section for stable nuclei, the in the separationenergy. total EMD cross section, o$MD( *2c+Pb), was estimated to be m35mb (Heckman and Lindstrom, 1976). If the projectile charge dependence of uEMDisassumed to be Zpr$, the enhancement of aEMD(**Li) compared with aEMDQ2C) is extremely large, by a factor of 80.
Nuclear Struchlre Studies
555
An EMD process is considered to occur due to the excitation of a projectile nucleus by the virtual photon field of a target nucleus. The EMD cross section is then calculated by,
(5.2) where IVYis the virtual photon spectrum and or is the photo-nuclear cross It is known that the electric giant dipole resonance (GDR) dominates section. or in high-energy collisions. It has been shown (Heckman and Lindstrom, 1976; Olson et al., 1981; Mercier et al., 1986) that the observed EMD cross section for stable nuclei can be well reproduced by this equation from the known photonuclear cross section, o+y+B+X), and theoretically calculated NY.
.
Soft El mode of excitation
Concerning the GDR of extremely neutron-rich nuclei, K. Ikeda (1992) suggested the possibility that the GDR could split into two components for nuclei with a neutron halo. One component would correspond to the oscillation of core protons against core neutrons with a normal frequency; the other correspond to the oscillation of the core against the skin and/or halo neutrons. Since the restoring force of the dipole oscillation is roughly proportional to the derivative of the density distribution, the frequency of the latter mode (denoted as the soft GDR) is expected to be very low. Within the framework of the GDR model, the energy of the soft GDR can be estimated. The GDR at normal excitation energies tEnorma GDR = 22 MeV estimated from the systematics) contributes to aIEMDonly by about 80 mb that agrees well with EMD cross section of 12C. Therefore, the main part of qmD is due to the excitation of the soft GDR. The integrated strength of the soft GDR was estimated to be 10% of the total El strength. (Since the El strength is roughly proportional to the square of the number of neutrons involved in each mode; the strength of the soft GDR : normal GDR = 22 : 62 for the llLi case.) In order to reproduce the observed 01EMDPLi+Pb), the energy of the soft GDR had to be E,ft GDR = 0.9+Os54.3 MeV if r,ft GDR was assumed to be 0.7 MeV. The energy (E,h GDR)was rather insensitive to the choice of r,,ft ~-J-JR. Thus, the excitation energy of the soft GDR turned out to be remarkably low. Qualitatively, the most essential conclusion drawn here is that the observed large EMD cross section, together with the rapidly decreasing virtual-photon spectrum, requires an appreciable strength of the El strength (resonance and/or continuum) to be located at a very low excitation energy, irrespective of the detailed structure.
5.56
I. Tanihata
Bertsch and Foxwell (1990) extended the Hartree-Fock model to calculate the El strength for IlLi. (Remember that his model explains the neutron halo.) The Random Phase Approximation (RPA) was used to calculate the dipole strength. The calculation shows a softening of the dipole mode, as shown in Fig. 5.3. However, their model predicts a smaller strength than needed to explain the observed enhancement in the EMD cross section. The calculated cross section is only about the 30% of the observed value. Even including the uncertainty in the experimental determination of the EMD cross section, more enhancement is needed. It is, therefore, suggested that any theory has to deal with the correlation between halo neutrons. Suzuki and To&a (1990) extended the cluster-orbital shell-model calculation for soft dipole excitation in the 9Li + n + n system. They found that quite a large strength (10% of the energy-weighted sum rule, or 30% of the non-energyweighted sum rule) is located at very low energy. They also obtained the EMD cross section to be 600 mb that agrees reasonably well with observed cross section. I “U
’
Dipole
1
’
1
I
‘2
Response
Free RPA
.... o.of
0
’
’
’
’
’ 10
’
m ’
Energy
’
’ 20
’
’
‘.......__) B -l 30
’
(MeV)
Fig. 5.3 Electric dipole strength distribution of 1*Li calculated by a RPA calculation.
Experimentally, an excited state of IlLi, as a possible candidate from the soft GDR, was observed at the excitation energy E, =1.2 fO.l MeV, from an experiment on double-charge-exchange (DCX) reaction induced by pions, llB( It-, rr+)“Li, at T, =164 MeV by Kobayashi et al. (1992). The spin and the parity of the state were assigned as being l/2+, 3/2+ or 5/2+ from the measured angular distribution, so that the state can be excited by the El transition.
557 NuclearStructureStudies However, it was not clear whether this state is the main component of the soft GDR or additional components are necessary, especially since the DCX reaction does not favor to excite a real collective state. Recently, the excitation distribution of l*Li EMD was measured by two groups: one at MSU (Ieki et al., 1993) and the other at RIKEN. (Shimoura et al., 1994) Also, the excitation distribution of llBe was measured in RIKEN (Nakamura et al., 1994). The El strength distribution in llBe has a special meaning because it has a single neutron halo, and thus no confusion due to neutron correlations arises. Also, it is known that a strong El amplitude exists between the first excited state f&=0.3198 MeV, In =1/2-j. It therefore shows that the expected strength of the low energy El resonance is in the bound state, and therefore the El strength observed in the EMD process is most likely due only to the nonresonant continuum. Figure 5.4 shows the El strength distribution. The direct breakup model calculation which give the El distributionas,
= gexp(21cr,) l+mo
Fig. 5.4
xez n*~(
2
* @(E,
0A
- E$‘* E,’
,
El strength
distribution
deduced
from the EMD cross section of 11Be.
2
1
&
3
[MeVl
(5.1)
558
I. Tanibata
is shown by the solid curve. Here, S is the spectroscopic factor for the 291/z state, K =@i?y/n , p is the reduced mass, Es is the separation energy, and QJ is the radius of the square-well potential \ relevant to the halo P neutron. The curve shows a good over-all fit to the data, thus indicating that x the transition is Elb 0. dominated and nona resonant. The angular distribution of the *oBe fragment in the recoil *lBe frame is shown in Fig. 5.5. The angular distribution is Fig. 5.5 Angular distribution of neutron emission by the EMD well reproduced by a shape process of the 1IBe + Pb -> l%e +n +X reaction. The angle confirming the dominance (+ is measured from the reaction plane. The angular distribution is consistent with a calculation based on the El transition. of El transition.
In MSU and also in RIKEN, two neutrons in coincidence with the 9Li fragment from *lLi + I% -> 9Li + n + n + X reaction wre measured. The upper part of Fig. 5.6 shows the obtained excitation-energy distribution by Ieki et al. (1993). The solid line in the figure shows a model calculation based on the BrightWigner distribution of a resonance at E&l.7 MeV with a width lYo=O.&O MeV. The dotted line in the figure is a prediction based on the direct break-up process by a Coulomb interaction. It shows that a better fit is obtained by the resonance picture as far as the excitation distribution is concerned. However, at the same time, they observed a systematic difference in the velocity between the neutrons and the 9Li fragment. Figure 5.6-c shows a velocity difference spectrum. The lithium-9 fragment has larger velocity than They claimed that this is due to a post do neutrons on the average. acceleration of the 9Li after the *lLi break up. From the observed difference in the velocity, llLi should breakup into 9Li and neutrons at a distance very close to the target. The resonance has its own lifetime, and thus decays much later after its production. They, therefore, concluded that this velocity difference is an evidence of a direct breakup.
559
Nuclear StructureStudies
However, recent theoretical studies by Bertsch and Bertulani (1993) show that the breakup is instant, and occurs at close distance independent of the mechanism, resonance or direct breakup. So far, several experiments seem likely to give controversial results for our understanding of the origin of the soft El strength in IlLi. Understanding the dipole strength is also important from the view point of correlations of halo neutrons. Further accurate measurements are thus awaited.
0
1.00
0.50
1.50
Ed (decay energy) [MeV] 400 200 0 3 0 !I 3100 50 0
-0.04
-0.02
0
0.02
0.04
V/C Fig. 5.6 Spectra of the 1lLi EMD process at MSU. (a) Decay energy spectrum of the 9Li+n+n system. (b) Spectrum of the longitudinal component of the center of mass velocity of the 9Li+n+n system, (c) Spectrum of the longitudinal component of the relative velocity VPV2n. The histogram shows the result of a Monte Carlo simulation without Coulomb acceleration.
6.
ELASTIC
AND
INELASTIC
SCATTERING
OF RADIOACTIVE
NUCLEI 6.1
Elastic scatterinv of 11h
A recent development in the intensity of radioactive beams has enabled us to study the elastic and inelastic scatterings of neutron-rich nuclei on proton target. Because a beam is the nucleus of interest, the reaction has to be studied by inverse kinematics. In an inverse kinematic measurement, the energy
560
I. Tan&a
resolution is not determined by the energy broadening of secondary beams, but by the energy resolution of the detected proton. As long as we use a thin target we can obtain a good energy resolution that is independent of the secondary beam resolution. Elastic scatterings were measured for 9Li + p, *lLi+p, *lLi+C, and llLi+Si at intermediate energies. Although a neutron halo has been observed in 11 Li from studies of high-energy collisions, an elastic scattering is expected to provide more detailed information about the structure of the halo. The result of the elastic-scattering angular distribution is shown in Fig. 6.1, together with the proton elastic scattering of other Li isotopes (6Li, 7Li, and 9Li) (Moon et al., 1992). The cross sections are similar for 6Li, 7Li, and 9Li; the absolute values are almost the same and the diffraction minimum systematically shifts to smaller angles for heavier isotopes, reflecting the increase in the matter radii. For l*Li, however, the cross flct..,.‘.,‘,‘....‘. .,.I “” section is smaller by 50% compared “‘“20 30 40 50 60 70 with that of other isotopes. An Center of mass angle (degree) optical potential fitting of the data was reported by two groups. In the Fig. 6.1 Proton elastic scattering angular distributions of Li isotopes. Although the cross first one, a global parameter set was sections show systematic behavior for 7Li, *Li, used as the starting point. Although and 9Li, sudden decrease is seen in 1 1Li this parameter set was determined spectrum. from reactions with mass numbers larger than 40 Warner et al., 1991), the parameter set without any modification fit the 6Li, 7Li, and 9Li data amazingly well as can be seen in Fig. 6.2-a, where an angular distribution is plotted as a ratio to the Rutherford scattering cross section However, it fails to reproduce the llLi cross section [the dashed curve in Fig.6.2-b]. Two different approaches were applied to reproduce the data. One is to change the parameters of the imaginary part of the potential. As shown by the dotted line (set A) in Fig. 6.2-b, a reasonable fit was obtained. The other is, contrarily, to change the real part of the potential. It also gives an equally good fit to the data, as indicated by the solid line (set B) in the figure. The largest difference between this potential is the depth of the real potential and the diffuseness of the imaginary potential. Set A has long tail in the imaginary part of the potential, while set B has a very shallow real potential. The shallow potential is considered to reflect the breakup process. Because of
Nuclear StructureStudies
561
the extremely small separation energy of the valence neutrons in **Li, the polarization potential is reflected as the repulsive force in the real potential. Because these two extreme selections of parameter sets can fit the data equally well, one can imagine that many other different sets of parameters can reproduce the data. Therefore, no unique optical potential can be obtained by this method. In a second paper, a more intuitive method was applied (Hirenzaki et al., 1993). Here, the optical potential was separated into two parts. One is the potential from the core (SLi), the other is that from the neutron halo. The optical potential of 9Li was first determined by fitting the 9Li + p data using the initial values of the parameter from the global potential. Then, an imaginary potential having the 100 same shape as the density distribution of the neutron halo determined from interaction cross section measurements, was added for *lLi. Then, the depth of the potential was treated as the fitting parameter. A reasonable fit to the data was obtained. With slight 5 adjustments of the other part of the potential, the fitting becomes 20 60 80 excellent. Therefore, a model that San4’ @es) explicitly includes the halo effect Fig. 6.2 Ratio of the elastic scatterings to the Rutherford also reproduces the data well. scattering cross sections. See text for the fitting by various models. Dot dashed curve is the results of Aleixo et al., (1991)
In the analysis presented above, however, it is not clear whether it is really necessary to take into account the halo structure in order to reproduce the data. In particular, it is not clear whether the effect of the breakup process is essential to the halo density shape or not. To answer this question, Suzuki et al., (1993) calculated the cross section based on the four-body model, (gLi+n+n)p, in order to take into account the halo structure. The Eikonal and adiabatic approximations were used to derive an optical potential that includes the breakup effect of halo neutrons to continuum states. They could reproduce the reduction of the cross section satisfactory. Figure 6.3 shows the optical
562
I. Tanhta
potential of 9Li and llLi thus obtained. The p-“Li optical potential has a much longer tail than that of gLi+p. They concluded that both the breakup effect of the halo neutron and the n-p exchange force are important in obtaining a good agreement with the da la. The elastic scatterings of IlLi were also measured with 12c (Kolata et al., 1992) and 28Si (Lewitowicz et al., 1992). As shown in Fig. 6.4, three-body calculations show fair agreement with the data if inelastic scatterings are taken into account (Thompson et al., 1993a). However, because of the uncertainty in the inelastic scattering, as well as the large error bars in the Fig. 6.3 Optical potential of I * Li and 9Li data, discrimination of the model could determined by the four-body model. not be made. This situation is similar to that for the %i case. Further experimental and theoretical studies are necessary to isolate the effect of the halo.
6.2
Elastic and Inelastic scatterinp of a&
We now consider the data of 8He at 73A MeV. Figure 6.5 shows an inclusive proton distribution from the 8He+p reaction as a function of the excitation energy of 8He at CICrnfrom 25 to 60 degrees (Korsheninnikov et al., 1993). One can see a strong peak for the ground state, reflecting the elastic scatterings. At around 3.7 MeV is a peak that corresponds to an inelastic scattering to a 8He When protons are detected in coincidence with forward *He, excited level. only the peak corresponding to the elastic scattering is seen. When protons are detected in coincidence with 6He, the elastic peak disappears and the peak at an excitation energy of 3.6 MeV dominates the spectrum. This state is also seen in several other coincidence conditions p+n, p+bHe+n, and p+7He, where 7He is identified by the invariant mass of the detected However, this level is not detected when a coincidence 6He+n system. includes 4He, as shown in Fig. 6.5-F. because both the 6He+n+n and 4He+4n channels are open, it indicates a small branching to 4He+4n, namely r4He+4n/rtotal<4%.
Nuclear StructureStudies
563
-
100
elastic only .. FaddeevM3 YabanaY1 (Op1/2)2 FaddecvQS FaddeevL6A Faddeev22 Core!potential only ..“.“....“_
8 a
_._I
---
”
-___
t
10-l 101
t
I+
elastic + inelastic - - - FaddeevQS FaddeewL6A - - - - Faddeev22 Core potential only
Fig. 6.4 Quasi-elastic scattering measurementof 1lLi +C and theoretical fits.
J.
564
I. Tanihata
3 80 60 40 20 0 -2
2
6
10 14 18
EIGHTMeV
-2 2
6
10 14 18
EIGHTMeV
Fig. 6.5 Inelastic scattering spectrum of 8He+p.
7.
STUDIES OF PARTICLE UNSTABLE NUCLEI OFF THE DRIP LINES
In addition to bound drip-line nuclei, studies of unbound states and unbound nuclei near the drip line are of great importance for understanding the Also, studies of nuclei beyond the drip line structure of drip-line nuclei. provide a unique opportunity to study nuclei under extreme conditions, such as an extreme Z/N ratio. By using the radioactive beams near the drip lines, one can now study unbound nuclei near to or even beyond the drip lines. Among the method are transfer reactions of the usual kind, such as (d,p) and (d?He) for neutron rich nuclei and the (p,d) and (d,n) reactions for the proton rich side, that provides direct accesses to unbound nuclei off the drip line. Also, invariant mass measurements can be efficiently used for producing unbound states and nuclei. Here, we see some recent studies by invariant mass measurements.
7.1
Invariant-mass
565
NuclearStructureStudies suectroscoDV
The development of high-intensity heavy-ion accelerators has enabled us to produce drip-line nuclei in large quantity and to use them as secondary beams. Interesting examples among them are unbound He isotopes (5He, ‘He, 9He, loHe) and excited states of bound He isotopes. In particular, loHe is interesting because it is a double magic-number nucleus beyond the neutron drip-line and has the largest IV/Z ratio ever observed. The invariant-mass measurement is a widely used method to search for particle resonances. In this method, several particles in the final state after the reaction are detected in coincidence, and the invariant-mass distribution is measured. The invariant mass is calculated from the observed four momenta (Pi) as,
where pi is the three dimensional momentum vector of the i-th particle. This method has two advantages: first, the mass (or excitation energy) is fully determined by the energy-momentum of particles in the final state so that the mass resolution is independent of the energy spread of the incident beam, and is thus suitable for secondary-beam application. Second, a thick target can be used because one detects only highenergy particles and thus has less effect from the energy loss in the target. For the same reason, it is more sensitive to a small production cross section.
140 120 100 $
go
8 s
60 40 2. 0 -2-l
0
12
3
4
5
6
J%EIe-n-n 9 MeV
Such experiments have recently been Fig. 7.1 Invariantm-s spectrumof *He +n+n system after the fragmentation of l ILi. The carried out at RIKEN using a magnetic peak at 1.2 MeV indicates the first spectrometer and neutron detectors observation of loHe. Lines show phase space (Korcheninnikov et al., 1994). A calculations. charged fragment and neutrons are detected in coincidence for the observation of unbound He isotopes. An invariant-mass spectrum of the *He +2n final state from the **Li fragment is shown as a function of the excitation energy from sHe +n+n (E*~H~~+J in Fig. 7.1, in which detector acceptance is corrected.
I. Tanibata
566
The spectrum is shown for a CDztarget. It shows a pronounced peak at 1.2 MeV with a small background at higher excitation energies. The phase-space calculation of the three-body final state including final state interactions are
loBe (l%, l%)lOHe, Elab=334.4 MeV 160 t...l.“‘““““““.-.‘-~
lOHe
z
8
2.7” < 81ab C 3.9” 8
80
20 a
40
60
80
100
120
Channel
Fig. 7.2 10~~ spectrum by a two-body reaction. In addition to tbe ground state, two excited states are suggested
shown by the curves in the figure. None of them including other phase space curves not shown here reproduce a peak near to the observed one, but
567
Nuclear Structure Studies
reproduce the shape at higher excitation. Therefore, the peak at 1.2 f 0.3 MeV is attributed to the *OHestate. Since the width of the spectrum is determined totally by the experimental resolution, they could determine the upper limit of the width as being I’c1.2 MeV. However, this width is already sufficiently small to reject the 2s1/2orbital for the valence neutrons. Helium-10 was also recently observed by nucleon-exchange two-body reactions (Ostrowski et al., (1994). Three levels are observed (see Fig. 7.2) and the lowest one (1.07 f 0.07 MeV above the 2n threshold) is consistent with the data by Korcheninnikov. Two other levels are located at excitation energies of 3.20 and 6.82 MeV from the lowest state.
Production of nuclei beyond the drip lines 1.
Invnriant ma58 spectroscopy ‘9-k: “Li -> %e+n+n 13Li : 14Be -> “ti +n+n ‘6Be . I78 -> ‘$e+ n+ n 18Be :. 19B
.. ... 2.
-> ‘$e+n+n+n+n
Transfer rpactiolrr (d, 3He) reaction
”
ti(d, 3He)‘%e, (d,p) reaction “Ii(d
‘%e(d,%e,‘h, .
... .
3. lhder reactions (p,d) lcactiom gc(PPk (d,n) wxtions ‘tJdsl)‘%,
4
. .. . ..
Invariant mass speetros~py 6Be -> a+p+p ‘ON -> 9c+tl
Fig. 7.3 List of the reactions for producing nuclei beyond the dripline. mass method is promising
Among them, the invariant
for many cases both beyond the proton and the neutron drip lines.
The invariant mass was also measured for the 6He +2n and 6He+n final states, and gives consistent energies of the excited levels with previous measurements based on other methods (Korsheninnikov et al., 1993). Invariant-mass spectroscopy is a powerful tool for studying unbound excited states and unbound nuclei. It is actually the best tool, at present, to study nuclei beyond the drip line, as demonstrated by the observation of loHe. It can also be used in the proton-rich side. Such a possibility is summarized in Fig. 7.3 for both limits
568
I. Tanihata
of the nuclear chart. Studies of 12J3Li, 15J6J7r18Be beyond the neutron drip line and Qe, IoN beyond the proton drip line would be possibile in the near future. Transfer reactions could also be used to study those nuclei beyond the drip line. Transfer reactions may provide more information than the invariantmass spectrum, if they are detected, because angular distributions can be studied in detail. One can also utilize the selectivity of the reaction by applying an appropriate reaction. In particular, (p,n) and (d,n) reactions have advantages in the proton-rich side because very low-energy neutrons can be detected without any energy loss problem in the target that determines the energy resolution for the detection of a charged-particle final state. Acknowledgment The author would like to express his thanks to Prof. T. Yamazaki for recommending that he write this article. He also gives thanks for all helps from Dr. D. Hirata and Ms. Odai for preparing figures and other materials.
Nuclear StructureStudies
569
References Alba, R. et al. (1993). Proc. of the Third Int. Conf. on Radioactive Nuclear Beams, Michigan State University, edited by D. J. Morrisey, Editions Frontiers, p31. Ann, R et al. (1993). Physics Letters m 55. Artukh, A. G. et al., (1993). Proc. of the Third Int. Conf. on Radioactive Nuclear Beams, Michigan State University, edited by D. J. Morrisey, Editions Frontiers, p45. Arnold, E, J. Bonn, A. Klein, R. Neugart, M. Neuroth, E. W. Otten, P. Lievens, H. Reich. W. Widdra. and ISOLDE Collaboration. (1992)Phys. I&. m 16. Asahi, K. et al., (1992). Hyperfine Interactions E 101. Bang, J. M, and I. J. Thompson. (1992) .Phys. Left. ~ 201-206. Becchetti, F. D. et al., (1991). NucI. In&r. and Meth. B56/57 554. Belmont, J. et al., (1992). Int. Workshop on “Physics and Techniques of Secondary Nuclear Beams” March 1992, Dourdan (France) Edited by J. F. Bruandet, B. Fermamdez, M. Bex, Editions Frontiers, 1992, p407. Bertsch. G. F, A. Brown, and H. Sagawa. (1989). Phys. Rev,C 39 1154. Bert&. G, and J. Foxwell. (1990). Phys. Rev w 1300. Bertsch, G. F, and H. Esbensen. (1991). Ann. Phys. 2pe 327. Bertsch, G. F, and C. A. Bertulani. (1993) . Nucl. Phys A556 136. Bibber, K. Van. et al. (1979) Phys. Rev. I&. g 840. Blank, B. et al. (1992)Z Phy. A343 375. Blank, B. et al. (1993). Nucf. Phys. m 408. Boyd, R. (1993). Physics Report 227 57. Brohm, T. et al. (1992). Nucl. Phys. m540. Buchmann et al., (1992). Proc. of the Second Int. Conf. on Radioactive Nuclear Beams, Louvain-la-Neuve 1991, edited by Delbar, Adam Hilger 1992, ~109. Conf. Proc. of the First Int. Conf. on Radioactive Nuclear Beams, Berkeley 1989, edited by W. D. Myers, J. M. Nit&e, and E. B. Norman, World Scientific 1990. Proc. of the Workshop on the Science of Intense Radioactive Ion Beams, Los Alamos, 1990, LA-11964-C UC-413. Proc. of Int. Symp. on Structure and Reactions of Unstable Nuclei, Niigata, Japan 1991, edited by K. Ikeda and Y. Suzuki, World Scientific. Proc. of the Int. Workshop on Unstable Nuclei in Astrophysics, Tokyo, Japan 1991, edited by S. Kubono and T. Kajino, WorId Scientific. Proc. of the Second Int. Conf. on Radioactive Nuclear Beams, Louvain-laNeuve 1991, edited by Delbar, Adam Hilger 1992. Proc. of the international Conference on Nucleus-Nucleus Collision IV, Edited by H. Toki, I. Tanihata, and H. Kamitsubo, Special issue of Nuclear Physics, North-Holland, Nucl. Phys. A538 (1992).
570
I.Tanilwa Proc. of the twelfth International Conference on Electromagnetic Isotope Separators and Techniques Related to their Applications, Edited by M. Fujioka, T. Shinozuka, and Y. Kawase, Nuclear Instruments and Methods in Physics Research B70 (1992). Proc. of the Sixth International Conference on Nuclei Far from Stability and the Ninth International Conference on Atomic Masses and Fundamental Constants, at Bernkastel-Kues, 1992, edited by R. Neugart and A. W&r, Institute of Physics Conference Series Number 132. Proc. of the Third Int. Conf. on Radioactive Nuclear Beams, Michigan State University, edited by D. J. Morrisey, Editions Frontiers (1993) Darquennes D. et al., (1990). Phys. Rev. u R804. DiGiacomo, N. J, R. M. DeVries, and J. C. Peng;. (1980). Phys. Rev. L&u. DiGiacomo, N. J, J. C. Peng, and R. M. DeVries. (1980). Phys. L&f. m 383. DiGiacomo, N. J, and R. M. DeVries. (1984). Comments Nucl. Part. Phys, u 111. Doubre, H. (1992). Int. Workshop on “Physics and Techniques of Secondary Nuclear Beams” March 1992, Dourdan (France) Edited by J. F. Bruandet, B. Fermamdez, M. Bex, Editions Frontiers, 1992, ~391. Dufour, J. P. et al., (1986). Nucl. Instrum. Methods in Phys. Res. &j_ 34. Fukuda, M. et al. (1991).Phys. L&t. B m 339. Fukunishi, N. T. Otsuka, and I. Tanihata. (1993).Phys. Rev. C #& 1648. Geissel,H. et al., (1992), Nucl. Instrum. Methods in Phys. Res. Bzp 286. Geissel, H. et al., (1993), Proc. of the Third Int. Conf. on Radioactive Nuclear Beams, Michigan State University, edited by D. J. Morrisey, Editions Frontiers (1993), Gelbke, C. K. et al. (1976). Phys. Rev. L.&t.37 1191. Goldhaber, A. S. (1974). Phys. L&t. a 306. Goldhaber, A. S, and H. H. Heckman. (1978). Part. Sci. Ann. Rev. Nucl. & Greiner D. E. et al. (1975). Phys. Rev. Letters, s 152. Guillemaud-Mueller D. et al. (19901, Phys. Rev. ,u 937. Ikeda, K. (1992). Nucl. Phys. M 355c. Harvey, B. G. (1990). Phys. I&t. j3238 11. Hayes, A. C. (1991). Phys. I&t. m 15, Hayes, A, C, and D. Strottman. (1991). Phys. Rev. a 2248. Heckman, H, and P.J. Lindstrom. (1976). Phys. Rev. L&t. 3756. Heckman, H. H, D. E. Greiner, P. J. Lindostrom, and H. Shwe. (1978). Phys. Rm. 171735. Hill, J. et al. (1988). Phys. Rev. -1722. Hill, J. et al. (1988). Phys. Rev. Lett. @_999. Hirata, D, H. Toki, T. Watabe, I. Tanihata, and B. V. Carlson. (1991). Phys. Rev. c &1467. Hirenzaki, S, H. Toki, and I. Tanihata. (1993). Nucl. Phys L9352 57. Hiifner, J, and M. C. Nemes. (1981). Phys. Rev. a 2538. Ieki, K. et al. (1993). Phys. Rev. Left. Ze 730. Igo, G. et al. (1979). Phys. L&t. m 151.
Nuclear Smcaue Studies
571
Kelley, J. H. et al. (1993). Proceedings of the ‘Third International Conference on Radioactive Nuclear Beams” East Lansing, Michigan, USA May 1993, edited by D. J. Morrissey, page345. Kobayashi, T, 0. Yamakawa, K. Omata, K. Sugimoto, T. Shimoda, N. Takahashi, and I. Tanihata. (1988). Phys. Rev. L.&t. de 2599. Kobayashi, T. et al. (1989) . Phys. I&t. u 51. Kobayashi, T. (1989a). Proceedings of the “First International Conference Radioactive Nuclear Beams” Berkeley, Ca 1989 Edited by W. D. Myers, J. M. Nits&&e, E. 8. Norman, World Scientific 1989, page 325. Kobayashi, T. (1992). Nucl. Phys. A538 343c. Kobayashi, T. (1993). Proceedings of the “Third International Conference on Radioactive Nuclear Beams” East Lansing, Michigan, USA May 1993, edited by D. J. Morrissey page169. T. Kobayashi et al. to be published. Kolata, J. J. et al. (1992). Phys. Rar. I&t.& 2631. Korsheninnikov, A. A. et al. (1993) . Phys. Z.&t.m 38. Korsheninnikov, A. A. et al. (1994). Phys. Lett. B a 31. Kox, S. et al. (1987). Phys. Rev. Cm678. Krasznahorkay, A. et al. (1991). Phys. Rev. Lett. 661287. Kubo, T. et al., (1992). Nucl. Instrum. Methods in Phys. Res. m 309. Lewitowicz, M. et al. (1993) . Nucl. Phys. A562 301. Lombard, R. J. (1990). 1. Phys. G 16 1311. Mercier , M. et al. (1986). Phys. Rev. m1655. Migneco et al., (1992). Int. Workshop on “Physics and Techniques of Seconda’ry Nuclear Beams” March 1992, Dourdan (France) Edited by J. F. Bruandet, B. Fermamdez, M. Bex, Editions Frontiers, 1992, p341. Minamisono, T , T. Ohtsubo, I. Minami, S. Fukuda, A. Kitagawa, M. Fukuda, K. Matsuta, Y. Nojiri, S. Takeda, and H. Kitagawa. (1992) . Phys. Rev. Mt. @ 2058. Mittig, W. et al. (1987). Phys. Rev. Mt. al889. Moon, C. -B. et al. (1992). Phys. L&m, B 297 39. Mueller A. C. and Anne R., (1990). Nucl. Instrum. Methods in Phys. Res. B56/57 559. Mueller, A. C, and B. M. Sherrill. (1993). Ann. Rev. AM. and Part. Sci. Vol. 43, Humbert F., et al., to be published. Myers, W. D. and W. J. Swiatecki. (1990). Annals of Physics m 401, and references therein. Nakamura, T. et al. (1994). Phys. Mt. B w 296. Nomura, T. 0992). Nucl. Intr. and Meth. in Phys. Res. m 407. Ogawa, Y, K, Yabana, and Y. Suzuki. (1992). Nucl. Phys. A543 722. Olson, D. et al. (1981). Phys. Rev. u 1529. Olsen, D. K. et al., (1992). Proc. of the Second Int. Conf. on Radioactive Nuclear Beams, Louvain-la-Neuve 1991, edited by Delbar, Adam Hilger 1992. ~131. Orr, N. A. et al. (1992). Phys. Rev. L&t. 69 2050. O&row&i, A. N. et al., (1994) Phys. Left. m 13.
512
I. Tanibata
Ozawa, A. et al. (1994). Phys. Lett. B m 18. Price, I?. B. et al. (1988). Phys. Rev. I_&. a 2193. Riisager, K. et al. (1992). Nucl. Phys. A54Q 365. Riisager, K. (1993). Proceedings of the “Third Znternational Conference on Radioactive Nuclear Beams” East Lansing, Michigan, USA May 1993, edited by D. J. Morrissey page281. Saint-Laurent, M. G. et al. (1988).GANIL repart P88-10, Sherril B. M. et al., (1990). Proc. of the First Int. Conf. on Radioactive Nuclear Beams, Berkeley 1989, edited by W. D. Myers, J. M. Nitchke, and E. B. Norman, World Scientific 1990. ~72. Sherril B. M., Morrissey, D. J., Nolen Jr, J. A., Orr, N., and Winger, J. A. (1992). Nucl. Instrum. Methods in Phys. Res. m 309. Shimoda, T., Miyatake, H. and Morinobu S. (1992). Nucl Instr. and Meth. m 320. Shimoura, S. (1991). Proceedings of the Second International Conference on Radioactive Nuclear Beams, Louvain-la-Neuve, Belgium 1991, edited by Th. Delbar, Adam Hilger, p 215. Shimoura, S. et al. (1994). Submitted for publication. Smith R. J. et al., (1990). Nucl. Inst. Methods && 26. Smith R. J. et al., (1991). Phys. Rev. Q@ 761. Suzuki, Y. and Y. Tosaka. (1990). Nucl. Phys. 517 599. Suzuki, Y. (1992). Proceedings of the 6th France-Japanese Colloquium on Nuclear Structure and Interdisciplinary Topics, Saint-Ma10 (France) October 1992. Suzuki, Y. K. Yabana, and Y. Ogawa. (1993). Phys. Rev. C 47 1317. S. Nagamiya, Y. Shida, N. Tanihata, I, H. Hamagaki, 0. Hashimoto, Yoshikawa, K. Sugimoto, T. Kobayashi, D. E. Greiner, N. Takahashi, and Y. Nojiri. (1985) . Phys. I.&f. 160B 380, Tanihata, I, H. Hamagaki, 0. Hashimoto, Y. Shida, N. Yoshikawa, K. Sugimoto, 0. Yamakawa, T. Kobayashi, and N. Takahashi. (1985a) . Phys. Rev. Lett. B 2676. Tanihata I. et al., (1985b). Hyperfine Interactions, 21 251. Tanihata, I, T. Kobayashi, 0. Yamakawa, S. Shimoura, K. Ekuni, K. Sugimoto, N. Takahashi, T. Shimoda, and H. Sate. (1988). Phys. Lett. B 2p6 592. Tanihata, I. (1988). Nucl. Phys. && 795c. Tanihata, I. (1989). On the Possibility Use of Secondary Radioactive Beams. TREATZSE ON HEAVY-ION SCIENCE, Vol. 8 , Edited by D. A. Bromley (Plenum Publishing Corporation, 1989) p443. Tanihata, I. (1991). Nucl. Phys. m 275~. Tanihata, I, T. Kobayashi, T. Suzuki, K. Yoshida, S. Shimoura, K. Matsuta, T. Minamisono, W. Chriestie, D. Olson, and H. Wieman. (1992) . Phys. Mt. B 282 307. Tanihata, I, D. Hirata, T. Kobayashi, S. Shimoura, K. Sugimoto, and H. Toki. (1992a). Phys. L.&t.B 289,261.
Nuclear Suucuue Studies
513
Thompson, I. J. and Russian-Nordic-British Theory collaboration. (1993); Proc. of Third Int. Conf on Radioactive Nuclear Beams at MSU 1993, edited by D. J. Morrissey. ~199. Thompson, I. J, S. ALkhalili, J. A. Tostevin, and J. M. Bang. (1993a) . Physics. Rt?v.c92,1364. Tosaka, Y, and Y. Suzuki. (1990). Nucl. Phys. m 46. Vamer, R. L. W. J. Thompson, T. L. McAbee, E. J. Ludwig, and T. B. Clegg. (1991).Phys. Rep. 2pL57. Villari, A. C. C. et al. (1989). Proceedings of XXVII International Winter Meeting on Nuclear Physics (Bormio, Italy Jan. 1989)~74. Villa@ A. C. C. et al. (1991). Phys. iI&. B 268 345. Westfall G. D. et al. (1979),Phys. Rev. Letters, fi 1859. Whitten, C.A. , Jr.; High Energy Physics and Nuclear Structure, ed. D.F. Measday, and A.W. Thomas (North-Holland, Amsterdam, 19801,p419. Zhukov, M. V, B. V. Danilin, D . V. Fedorov, J. 6. Vaagen, F. A. Gareev, and J. Bang. (1991).Phys. L&t. B m 19. Zhukov, M. V. B, V. Danilin, D. V. Fedorov, J. M. Bang, I. J. Thompson, and J. S. Vaagen. (1993).Physics Reports a 151.