Paper

The electric dipole response of exotic nuclei

and

Published 28 January 2013 © 2013 The Royal Swedish Academy of Sciences
, , Citation T Aumann and T Nakamura 2013 Phys. Scr. 2013 014012 DOI 10.1088/0031-8949/2013/T152/014012

1402-4896/2013/T152/014012

Abstract

The electric dipole (E1) response of exotic nuclei, with a focus on the soft E1 excitation of halo nuclei and the pygmy dipole resonance (PDR) of neutron-skin nuclei, is reviewed here from an experimental point of view. Since the discovery of halo nuclei in the late 1980s, the particular dipole response of these loosely bound nuclei located at the neutron-drip line resulting in an extraordinary enhancement of the electric dipole strength at low excitation energies (soft E1 excitation) has been one of the phenomena under intense investigation both in experiments and in nuclear theories. We discuss how such phenomena have been recognized as a unique feature of halo nuclei and how it is applied to probe the halo structure in a quantitative way. Another intriguing feature of the E1 response of neutron-rich nuclei is the PDR in neutron-skin nuclei, which is viewed as a dipole vibration of excess neutrons against a core nucleus and which appears in the low-energy part of the E1 strength function below the giant dipole resonance. The status of the experimental evidence for such a redistribution of dipole strength for neutron-rich nuclei compared to stable ones is reviewed. The relation of the dipole response to the symmetry energy in the equation of state of asymmetric nuclear matter and to the neutron-skin thickness is also briefly discussed. Finally, we outline some perspectives for future experimental investigations toward heavier and more neutron-rich systems. The experimental results reviewed here have been accumulated over the last two decades at fragmentation facilities worldwide at GANIL, GSI, MSU and RIKEN.

Export citation and abstract BibTeX RIS

1. Introduction

When we add neutrons one by one to a β-stable nucleus, how does the nucleus change its structure and dynamics? This is one of the fundamental questions that have been raised in nuclear physics. Indeed, a large excess of the neutron number (N) over the proton number (Z) in a nucleus induces anomalous and intriguing features. Figure 1 illustrates such a drastic change in the structure toward the neutron-drip line. A β stable nucleus, as shown in figure 1(a), has almost identical Sn (one neutron separation energy) and Sp (one proton separation energy). The stable nucleus also has a similar neutron and proton density profile, i.e. the root mean square (rms) radius of the proton distribution is almost identical to that of the neutron's. This situation drastically changes as the N/Z ratio becomes larger. As shown schematically in figure 1(b), a larger N/Z ratio causes the difference of Fermi energies between protons and neutrons, Sp − Sn, and accordingly a thick neutron skin of 0.2–0.9 fm develops [1, 2]. Note that the difference of the Fermi energies reaches almost 20 MeV near the drip line, which is almost half the depth of the mean field potential of a normal nucleus. At the neutron-drip line (c), when Sn itself approaches zero, the neutron wave function extends far outside of the nuclear mean field as a result of quantum tunneling and a neutron halo is formed [35]. In this case, the value of Sn is typically less than 1 MeV, much smaller than for normal nuclei of about 8 MeV. Further interesting features of neutron halo and skin nuclei are the resultant dynamical aspects as revealed, for instance, in the electric dipole response as discussed here.

Figure 1.

Figure 1. The change of the mean field potential (top) and density profiles (bottom) from β stable (a) to neutron-rich nuclei (b) and to much more neutron-rich nuclei near the neutron-drip line (c) is schematically shown. The upper panels indicate how the change of the ratio of proton/neutron numbers can induce a significant difference of Fermi energies between protons and neutrons as measured by Sp − Sn. Such a difference causes the formation of a thick neutron skin as in the bottom panel of (b). When Sn approaches zero, the neutron halo structure appears as a consequence of the quantum tunneling.

Standard image

The study of the physics of halo nuclei was initiated in the mid-1980s by Tanihata et al [3, 6, 7] when they started to apply high-energy fragmentation reactions to produce very exotic nuclei by using heavy ions at about 800 MeV per nucleon at the Bevalac at LBNL (Lawlence Berkeley National Laboratory). They measured systematically the interaction cross sections of light neutron-rich nuclei up to the neutron-drip line from He to Be isotopes and found that the most neutron-rich Li isotope, 11Li, exhibits by far a much larger matter radius as compared to the conventional r0A1/3 systematics [3]. While it was not evident a priori if the origin would be a strong deformation or a long tail in the matter distribution [6], additional experimental information on the magnetic moment obtained at ISOLDE [8] (and later also of the quadrupole moment [9]) suggested an explanation as detailed by Hansen and Jonson [4] that the large interaction radius is related to a long tail of the neutron wave function due to the small binding, which they named 'Halo'. In the same paper, Hansen and Jonson also predicted that, as a consequence, the Coulomb breakup cross section of such halo nuclei should be extremely large. Direct experimental evidence of the halo structure of 11Li, where two valence neutrons are extended over the densely packed 9Li core, then came from the observation of the narrow momentum distribution of 9Li, following the breakup of a 11Li projectile with a carbon target, by Kobayashi et al [10] (transverse component of the momentum), and later confirmed by Orr et al [11] (longitudinal component of the momentum). These observations directly reflect the small internal momentum equivalent to the extended density distribution of the two valence neutrons, which are removed in the reaction. The third important observation on 11Li was the anomalously large electromagnetic dissociation (EMD) cross section of close to one barn by Kobayashi et al [5]. They bombarded the heavy Pb target with a 11Li beam and observed inclusively the 9Li fragment [5]. This was interpreted as evidence of an enhanced electric dipole (E1) response at low excitation energies, called 'Soft E1 excitation' as predicted by Hansen and Jonson [4]. These three observations, large interaction cross sections, narrow momentum distribution of the fragment and the enhanced electric dipole response at low excitation energies, have been the major signals of any halo system. The pioneering experiments by Tanihata et al [3], which led to the discovery of the halo nuclei, constitute the birth of physics using reactions with radioactive beams. Reaction experiments with radioactive ion (RI) beams have been extremely successful over the last two decades; in conjunction with increased efforts in nuclear theory, they have allowed to unravel and understand the properties of these exotic nuclear systems.

In this article, we focus on one observable to investigate exotic nuclei, that is, the electric dipole (E1) response. Considering the E1 response of stable nuclei, where the giant dipole resonance (GDR) around Ex ∼  80A−1/3 MeV (∼13–20 MeV) exhausts most of the E1 strength, the significant E1 strength well below the GDR energy for neutron halo and neutron-skin nuclei is very unique. Hence, the E1 response of exotic nuclei, in particular that of neutron-rich nuclei exhibiting neutron-halo or neutron-skin structure, has been a major subject of physics of exotic nuclei. Experimentally, the E1 strength function of exotic nuclei can be accessed by using heavy-ion-induced electromagnetic excitation. Most of the experiments discussed here study the neutron or particle decay after excitation. In this case, the process is called the Coulomb breakup4.

The mechanism of the enhanced E1 response for halo nuclei has been a major issue of the physics of halo nuclei. Hansen and Jonson [4] discussed the possibility of 'soft electric dipole mode' for 11Li, which is considered as having a non-resonant character, inferred from the analogy of deuteron. Meanwhile, as shown in figure 2, Ikeda attributed the enhanced low-energy E1 strength to the soft dipole resonance (SDR), caused by the low-frequency motion of the charged core nucleus against the low-density neutron halo [12].

Figure 2.

Figure 2. The upper panel shows the vibration of the 'neutron fluid' relative to the 'saturated core'. The bottom panel shows the transition probability as a function of excitation energy. The lower-energy bump corresponds to a 'new soft-dipole giant resonance', whereas the higher-energy bump corresponds to 'GDR analogous to the conventional one'. The figure is reprinted with kind permission from Ikeda [12].

Standard image

As shown in detail in section 3, the primary mechanism is now established as having a non-resonant character, called the direct breakup mechanism, as has been investigated for one-neutron halo nuclei such as 11Be [1315], 15C [16, 17] and 19C [18]. The cases of two-neutron halo systems, such as 6He [19] and 11Li [20], are much more complicated. The current understanding for two-neutron halo cases is also non-resonant origin, but the initial two-neutron correlation and the final state interaction (FSI) affect largely the E1 spectra. From the non-resonant feature of this excitation, we may call this phenomena 'soft E1 excitation'. Soft E1 excitation has become one of the sensitive tools to probe the microscopic structure of halo states, such as the single particle state for the one-neutron halo nuclei. In the case of two-neutron halo nuclei, the dipole strength at low energy reflects the intrinsic correlations of the halo neutrons, as well. This has been studied quantitatively so far only for the cases of 6He [19] and 11Li [20], which will be discussed in section 3.2.

For neutron-rich systems not exhibiting a halo structure, the dipole response is expected to change as well. In contrast to stable nuclei for which the dipole response is dominated by the GDR, new collective modes have been suggested on the basis of macroscopic models. In 1971, Mohan et al [21] discussed the possibility of a new excitation mode related to excess neutrons in a three-fluid model. Experiments with stable nuclei, in particular neutron capture measurements, reported low-lying dipole strength below the threshold at around 7 MeV, which has been discussed in this context and has often been called pygmy resonances (for an overview on the experimental status of pygmy resonances in stable nuclei, see the recent review article by Savran et al [22]). The soft dipole mode for halo nuclei discussed by Hansen and Jonson [4] has triggered new interest and the hope that such a new mode could be observable in neutron-rich nuclei in general. Suzuki et al [23] have studied this possibility in the framework of the Steinwedel and Jensen hydrodynamical model and predicted the strength and position of a such a mode as a function of neutron excess and mass. The situation of a vibration of excess neutrons against a core is illustrated schematically in figure 2, which was originally proposed as SDR for neutron halo nuclei.

While the dominant mode of excitation in stable nuclei is the collective out-phase vibration of neutrons against protons as the GDR located at around 13–20 MeV, the vibration of less bound excess neutrons against the core is expected at a lower frequency due to the weaker restoring force.

Such a redistribution of the dipole strength for neutron–proton asymmetric nuclei and the appearance of a soft vibration mode has been predicted also by microscopic calculations. Figure 3 shows as an example a result from the work of Vretenar et al [24] for nickel isotopes.

Figure 3.

Figure 3. Left frames: dipole strength distribution calculated in relativistic RPA for nickel isotopes [24]. Right frames: transition densities for the high-energy peak (GDR) and the low-energy peak (PDR). Figure reprinted from Vretenar et al [24]. Copyright (2001), with permission from Elsevier.

Standard image

The left frames show the strength functions calculated in a relativistic random phase approximation (RPA) approach. For the neutron-rich isotopes 68,78Ni a peak-like structure at around 9 MeV excitation energy is visible, well below the GDR region around 17 MeV. The transition densities for the high-energy peaks and the low-energy peak show pronounced differences. While the high-energy transitions exhibit the typical out-of-phase oscillation of neutrons versus protons, a pure neutron oscillation is visible at the surface region for the low-energy peak, which corresponds to the vibration of the neutrons against an N = Z core as discussed in the macroscopic picture. A comprehensive discussion from a theoretical point of view can be found in the review article by Paar et al [25].

The first experimental study of the full E1 response of neutron-rich nuclei has been carried out at GSI for the oxygen isotopic chain [26]. It was found that the dipole response indeed changes dramatically with adding more and more neutrons. A clear separation into two energy domains representing the two modes of vibration discussed above has not been observed. A later measurement for the neutron-rich tin isotopes 130,132Sn, however, gave clear evidence for a low-lying pygmy dipole resonance (PDR) located around 10 MeV excitation energy [27].

Since the change in dipole response for neutron-rich nuclei is caused by the less bound excess neutrons, which are responsible for the development of a neutron skin as well, it is clear that the low-lying dipole strength is related to the neutron-skin thickness. It has been noted that the predicted dipole strength function depends sensitively on the parameters of the effective interactions used, which describe the isospin-asymmetric part of the nuclear equation of state. In turn, it was suggested that such parameters may be constrained from measurements of the pygmy dipole mode in exotic nuclei [28] or from the dipole strength or polarizability in general [29]. First attempts in this direction have been made by using data on neutron-rich nuclei [30, 31] and also for the stable nucleus 208Pb [32].

In this article, we first give in section 2 a brief introduction to the method of Coulomb excitation at intermediate and high energies. We then discuss the soft E1 excitation of halo nuclei and its spectroscopic relevance in section 3. Here, we will emphasize also the case of two-neutron haloes, where the dipole response gives insights into correlations among the halo neutrons. Section 4 is devoted to the dipole response of non-halo neutron-rich nuclei and the PDR, including a discussion of its relevance for constraining the density dependence of the symmetry energy and its relation to the neutron-skin thickness. Finally, section 5 gives a brief summary and points out the open questions and future perspectives of this field in the era of the new-generation high-energy RI beam facilities such as FAIR at GSI, FRIB at MSU and RIBF (RI Beam Factory) at RIKEN.

2. Experimental method

2.1. Coulomb excitation and breakup at intermediate and high energies

Coulomb excitation is a process where a fast projectile at relativistic energies is excited by an impulse of the Lorentz-contracted Coulomb field of a high-Z target. When the final state is above the particle decay threshold and thus breaks up, this specific channel is called Coulomb breakup or EMD, as schematically shown in figure 4.

Figure 4.

Figure 4. The Coulomb breakup process is shown schematically for a projectile with AZ on a Pb target, decaying into A−1Z + neutron. The excitation process can be equivalent to the absorption of a virtual photon with Eγ(=Ex).

Standard image

Coulomb excitation at low incident energies below the Coulomb barrier has long been used successfully in spectroscopy of low-lying excited states of stable nuclei, where the nuclear excitation does not contribute. More recently, however, it has been proven that the Coulomb excitation at relativistic energies has advantages over low-energy Coulomb excitation because of the large experimental yield due to the kinematic focusing and the availability of a thick target of the order 0.1–1 g cm−2. It should be noted that the nuclear contribution can be estimated by using a less electromagnetic sensitive target (such as 9Be and 12C). The nuclear breakup contribution can also be controlled by the selection of the scattering angle, or equivalently of the impact parameter. Since the intensity of the RI beam is generally low, the Coulomb excitation at intermediate/high energies becomes a powerful spectroscopic tool. This also matches the intermediate-/high-energy RI beams available at the heavy-ion facilities based on the fragmentation reactions, such as GANIL, GSI, MSU and RIKEN.

The Coulomb excitation at relativistic energies can be expressed as a photo-absorption process induced by a virtual photon. The way of treating Coulomb excitation/breakup in terms of the virtual photon is called equivalent photon method. In this method, the Coulomb excitation/breakup cross section for the electric excitation with a multipolarity λ can be described as

Equation (1)

where NEλ(Ex) is the number of virtual photons of the electric multipole λ with photon energy Ex [33]. Hence, the reduced transition probability B(Eλ), which contains the physics essence of Eλ response, can be directly extracted from the excitation (breakup) cross section spectrum. The simplicity with which the cross section can be factorized in this way is one of the advantages of using the Coulomb excitation/breakup. The magnetic excitation is treated similarly, and the corresponding photon number can be seen, for instance, in [33].

The spectrum of the virtual photon can be calculated since it is an electromagnetic process [33]. In figure 5, the virtual photon spectrum for E1 excitation is shown for a projectile at 70, 250 and 800 MeV per nucleon on a Pb target. The impact parameter cut is fixed at 12.3 fm here, corresponding to the halo nucleus 11Be on Pb [13]. The characteristic feature of this photon spectrum is that it is a nearly exponentially falling function, so that the low-energy excitation is, in particular, amplified. It is interesting to note that the curves are crossing around Eγ ∼ 7 MeV. The yield at the low excitation energies is thus larger for the lower incident energies, while it is the inverse for the higher incident energies. Hence, the soft E1 excitation has a larger cross section for the lower incident energies, while it is almost similar or a little larger for the higher incident energies in exciting PDR at around Ex = 8–10 MeV. The GDR is excited more favorably for the higher incident energies. The incident energy dependences of the cross sections for the different multipolarities and for different excitation energies are shown in figure 6.

Figure 5.

Figure 5. E1 virtual photon spectrum for a projectile at 70, 250 and 800 MeV per nucleon with a Pb target for an impact parameter cut of 12.3 fm. It is shown as a function of the photon energy Eγ(=Ex).

Standard image
Figure 6.

Figure 6. Coulomb breakup (excitation) cross sections estimated for the E1, E2 and M1 transitions, with one Weisskopf unit (W.u.) for a nucleus with A = 11, are shown as a function of the incident energy. A strong dependence on the multipolarity and the incident energy is seen, which could be useful in distinguishing the multipolarity.

Standard image

2.2. Inclusive measurement and kinematically complete measurement

There are two major methods of measuring the Coulomb breakup, the inclusive method and the kinematically complete measurement (final-state exclusive measurement). The inclusive measurement extracts the 1n (2n) removal cross section (inclusive cross section) for the aimed nucleus. For instance, in the case of the 11Li breakup, the 9Li produced in the reaction of a 11Li projectile with the target is counted as in [5]. This corresponds to measure

Equation (2)

The cross section is now an integration of the product of NE1(Ex) and dB(E1)/dEx in terms of Ex from the decay threshold energy Eth to infinity. Since the photon spectrum falls nearly exponentially with Ex, as shown in figure 5, σ(E1) becomes large when the dipole strength B(E1) is concentrated at low excitation energies (soft E1 excitation) as in halo nuclei.

The advantage of the inclusive measurement is that the experiment is simpler, and the yield is higher since one does not need to measure neutron(s), compared with the kinematically complete measurement. The experiment is feasible even for secondary beam intensities of the order of 1 count per second or even less. However, as one cannot map the B(Eλ) distribution, the analysis covered in this method is restricted. Nevertheless, one can obtain evidence of the halo state as demonstrated for 31Ne [34], which is discussed in section 3.

The kinematically complete measurement maps the B(Eλ) distribution as a function of the relative energy Erel. It reconstructs the invariant mass from the four-momentum vectors of all the particles following the breakup of the projectile. As an example, we consider the case of AZ + Pb → A−1Z + n + X, namely the one neutron removal of AZ with a Pb target. One can then measure exclusively the final states of the excited AZ. In this case, the invariant mass M(AZ*) can be related to the four-momentum vectors of the outgoing particles as in

Equation (3)

where E(A−1Z) and E(n) are the total energy of the fragment (A−1Z) and that of the neutron, respectively. The relative energy Erel between the fragment and the neutron is then related to M(AZ*) as

Equation (4)

where M(A−1Z) and mn denote the mass of the fragment and of the neutron, respectively. The excitation energy Ex is related to the relative energy by Ex = Erel + Sn. Note that when the fragment is produced as an excited bound state, the mass of the fragment should be replaced by

Equation (5)

where the second term is the summation of the γ-ray energies associated with the daughter nucleus AZ when it is produced in a bound excited state. The excitation energy then becomes

Equation (6)

For the soft E1 excitation of halo nuclei, the contribution of the bound excited states is generally weak due to the rapidly falling function of the virtual photons. On the other hand, higher excitation as in PDR and GDR involves more excited daughter state contributions, where γ-ray energy measurements are essential. Typical setups for the kinematical complete measurement are shown in figures 8 and 17 for the RIKEN setup and in figure 21 for the GSI setup, which are explained in later sections.

3. Soft E1 excitation of halo nuclei

The E1 response of halo nuclei has a unique feature that the excitation also occurs strongly at low excitation energies at Ex ∼ 1 MeV, much below the GDR peak. In this section, we first clarify the underlying mechanism. We then discuss how this excitation can be related to the halo property and we show how this can be applied to the spectroscopy. Figure 7 shows the location of neutron halo nuclei, whose E1 response is studied by either the kinematically complete measurement or inclusive Coulomb breakup cross section. At RIBF at RIKEN, other halo nuclei such as 19B and 22C have been studied by kinematically complete measurements, and some heavier neutron-rich nuclei such as 37Mg have been studied inclusively, whose analysis is in progress. Hence, in the near future, the E1 response will be clarified in those nuclei as well.

Figure 7.

Figure 7. The halo nuclei and the current status of the Coulomb breakup experiments of those nuclei are shown in the nuclear chart up to Z = 20 (Ca). The known neutron halo nuclei (published) are shown as indicated. The neutron halo nuclei studied by a kinematically complete measurement (published) are shown by filled circles, and the ones studied only by inclusive Coulomb breakup (published) are shown by the gradation of filled circles.

Standard image

A one-neutron halo nucleus has a simpler structure due to its two-body nature. Unlike a two-neutron halo nucleus, which has a three-body Borromean structure, it does not suffer from strong FSIs in the breakup reactions. Hence, we first describe the case of one-neutron halo nuclei and discuss the mechanism causing the soft E1 excitation.

3.1. Soft E1 excitation of a one-neutron halo nucleus

First we clarify the underlying mechanism of soft E1 excitation and its spectroscopic significance by taking the example of 11Be. Then we show the application to clarify the structure of 19C, which had been controversial before the Coulomb breakup experiment [18], and to the case of 15C, which is important in the stellar reaction.

3.1.1. Soft E1 excitation of 11Be.

11Be has been a benchmark halo nucleus since it has a simple single-neutron halo structure 10Be + n with a well-determined separation energy Sn = 504(4) keV. The ground-state structure is known to be dominated by the following two configurations:

Equation (7)

where the first term corresponds to the halo state. There were spectroscopic studies of 11Be by using the transfer reaction,10Be(d,p)11Be, from which the dominance of 1s1/2 neutron in the ground state (α2 = 0.73 [35] and 0.77 [36]) was extracted. More recently, the knockout reaction of 11Be [37] and the 10Be(d,p) reaction in the inverse kinematics [38] also showed similar values close to 0.7. We also show here that the Coulomb breakup can be used to extract the spectroscopic factor for the halo state.

The earlier Coulomb breakup experiments were devoted, on the other hand, primarily to the two-neutron halo nucleus 11Li [3941]. However, they failed to clarify the mechanism of the soft E1 excitation, where the obtained B(E1) strength distributions resulted in inconsistent spectra as shown later in figure 17 (right). Theoretical interpretation of the soft E1 excitation of 11Li requires an understanding of the correlations of the two neutrons and other two-body FSIs. Hence, the Coulomb breakup of the simpler 1n halo nucleus 11Be played a significant role in understanding the underlying mechanism of the soft E1 excitation.

The first full kinematical complete measurement of 11Be was made at RIKEN at 72 MeV per nucleon [13], where the E1 spectrum was obtained, and evidence for the direct breakup mechanism of the Coulomb breakup is shown. There the spectroscopic factor α2 was also extracted and the spectroscopic significance was suggested for the first time. In GSI, at much higher energies of 520 MeV per nucleon [14], the Coulomb breakup of 11Be was measured to support the result of [13] with higher statistics. This experiment also measured γ-ray in coincidence to pin down the component of the second term of equation (7). It favors comparatively higher excitations due to the incident energy dependence of the virtual photons shown in figure 5. The third experiment performed at RIKEN [15], which has further higher statistics and better resolutions, utilized the angular distribution to refine the E1 spectrum. Here, we explain the essence of the Coulomb breakup and the soft E1 excitation of 11Be by showing the latter experiment of RIKEN [15].

The experimental setup used in this experiment [15] is shown in figure 8. The 11Be secondary beam, produced by fragmentation of an 18O beam at 100 MeV per nucleon at RIPS at RIKEN, bombarded a Pb target at an average energy of 68.7 MeV per nucleon. The momentum vectors of the beam ($\vec {P}(^{11}{\mathrm { Be}})$ ) as well as those of the outgoing 10Be fragment ($\vec {P}(^{10}{\mathrm { Be}})$ ) and the neutron ($\vec {P}(n)$ ) were measured in coincidence.

Figure 8.

Figure 8. Experimental setup used in the kinematically complete measurement of Coulomb breakup of 11Be at 69 MeV per nucleon at the RIPS facility at RIKEN. Figure reprinted with permission from Fukuda et al [15]. Copyright (2004) by the American Physical Society.

Standard image

The invariant mass of 11Be was then extracted to deduce the Coulomb breakup spectrum as a function of 10Be − n relative energy, Erel. Spectra of the breakup cross sections with the Pb and C targets are shown in figure 9 (left) (a) and figure 9 (left) (b), respectively. The breakup data on the carbon target, where nuclear breakup is dominant, are used to investigate the characteristic features of the target dependence. The cross sections are plotted for different selection of scattering angle θ in the center of mass of the projectile and the target.

Figure 9.

Figure 9. Left: breakup cross sections as a function of Erel of 10Be and the neutron in the breakup of 11Be + Pb at 69 MeV per nucleon (a) and 11Be + C at 67 MeV per nucleon (b). See text and [15] for details. Right: angular distribution of the 11Be in the center of mass of 11Be + Pb for the two Erel ranges shown. The grazing angles are indicated by θgr(=3.8°). Solid curves are calculations of the ECIS code, whereas the dotted ones are calculations of the equivalent photon method. Both calculations are E1 only, and no nuclear excitation is included. Figure reprinted with permission from Fukuda et al [15]. Copyright (2004) by the American Physical Society.

Standard image

The striking contrast is evident between the spectra for the two targets. The total one-neutron removal breakup cross section for the Pb target is 1790 ± 20(stat.) ± 110(syst.) mb, whereas that for the C target is 93.3 ± 0.8(stat.) +5.6−10.3(syst.) mb, about 1/20 of the former5. The large difference in cross sections can be attributed to the different reaction mechanisms between Pb and C targets. The former is dominated by Coulomb breakup, whereas the latter is dominated by the nuclear breakup. It should be noted that the ratio of the 'nuclear' breakup cross sections for these two targets would be about 2 [15], much smaller than this observed ratio.

The spectral shape exhibits a remarkable difference as well according to the target. The spectra for Pb for the two angular ranges have an asymmetric peak at very low Erel, whereas that for C for the whole angular range shows two prominent resonance peaks (Ex = 1.78 MeV, 5/2+ and Ex = 3.41 MeV, 3/2+) upon the structureless continuum. This also suggests the Coulomb breakup dominance for the Pb target data, and the nuclear breakup dominance for the C target data.

In this experiment, the scattering angle was measured and used effectively to control the reaction channel. The scattering angle is obtained by the difference between the incoming momentum vector of 11Be and the outgoing vector (sum of momentum vectors of 10Be and the neutron) in the center of mass (c.m.) frame of the 11Be + Pb(C). In the Coulomb breakup, the scattering particle (c.m. of 10Be + n) largely follows the Rutherford trajectory, the impact parameter b can be determined event by event from θ as b = a cot(θ/2), where a represents half the distance of closest approach in the classical Coulomb head-on collision. Although the nuclear breakup does not follow this rule, the selection of forward angles can be a good way of selecting the Coulomb breakup component. This can be seen in the forward-angle-selected spectra for Pb and C targets. There, the spectra have almost identical asymmetric shapes for both targets, which follow perfectly the calculation of the direct breakup mechanism of the Coulomb breakup of the one-neutron halo nucleus, as shown by the solid curves. The result shows that the direct breakup mechanism is indeed the underlying mechanism of the soft E1 excitation. It is also noted that for the Pb target, 6° data (whole acceptance) can also be explained by the direct breakup calculation for the most part. This again proves the dominance of the Coulomb breakup for the Pb target.

The selection of the most forward angles, well within the grazing angle (θgr = 3.8°), has been found useful in extracting almost purely the 'Coulomb' component of the breakup, which was confirmed by detailed reaction theories [4244]. For the whole acceptance data, we find a slight deviation from the direct breakup curve, which can be attributed to nuclear breakup and higher order Coulomb breakup effects.

The B(E1) distribution was extracted from the forward-angle-selected data on Pb. In this case, it is useful to use the double differential cross section formula for the E1 excitation in the equivalent photon method, written as

Equation (8)

The virtual photon number is extracted by integrating over the selected angles in this case. The B(E1) spectrum was obtained by dividing the dσ/dErel by this integrated photon number. The result is shown in figure 10, in agreement with the direct breakup mechanism shown in the solid curve. The integrated B(E1) obtained from the data selected for the forward angles amounts to 1.05 ± 0.06 e2 fm2 corresponding to 3.29 ± 0.19 W.u. (Weisskopf unit) for Ex ⩽4 MeV, which is huge, considering that the E1 strength in this energy region for normal nuclei is negligible (below 0.1 W.u.).

Figure 10.

Figure 10. B(E1) strength distribution for 11Be as a function of Ex, obtained from the angle-selected Coulomb breakup data on a Pb target (θ < 1.3°). The solid curve is the result of a calculation with the direct breakup mechanism with α2 = 0.72. Figure reprinted from Nakamura et al [13]. Copyright (1994), with permission from Elsevier.

Standard image

In the soft E1 excitation, the non-energy weighted E1 cluster sum rule is useful in extracting the geometrical information. For the one-neutron halo nucleus, the sum rule can be expressed as

Equation (9)

which shows that the integrated B(E1) is proportional to 〈r2〉, where r represents the distance of the halo neutron from the c.m. of the core. For 11Be, Fukuda et al [15] estimated $\sqrt {\langle \, r^2\,\rangle } = 5.77\pm 0.16\,{\mathrm { fm}}$ by summing the E1 strength up to Ex = 4 MeV, which is 1.05 ± 0.06 e2 fm2. Note, however, that the sum should be taken up to a sufficiently high excitation energy, and also take into account the transitions to bound final states [45, 46], which resulted in $\sqrt {\langle \, r^2\,\rangle } = 6.4 \pm 0.5\,{\mathrm { fm}}$  [47].

3.1.2. Direct breakup mechanism.

It was clearly shown above that the direct breakup mechanism explains the B(E1) spectrum of the one-neutron halo nucleus 11Be. The direct breakup itself has been known for long, for instance, for the photodisintegration of the deuteron [48].

In the direct breakup mechanism, the 1n-halo nucleus, in this case 11Be, breaks up into 10Be and a neutron without forming any intermediate resonances. The possibility of the SDR being a main mechanism of this large E1 strength is unlikely [1315].

The matrix element of the direct breakup is described as

Equation (10)

where $\Phi _{\mathrm {i}}(\vec {r})$ and $\Phi _{\mathrm {f}}(\vec {r},\vec {q})$ represent the wave function of the ground state and the final state in the continuum of the neutron relative to the core, respectively. $\vec {r}$ is the relative coordinate of the neutron relative to the c.m. of the core. $\Phi _{\mathrm {f}}(\vec {r},\vec {q})$ is also a function of the relative momentum $q=\sqrt {2\mu E_{\mathrm { rel}}/\hbar }$ . The electric dipole operator is $\hat {T}(E{\mathrm { 1}})(=rY^{(1)}(\Omega)$ ).

It is worth considering a simple formula for the case where the ground state is a pure single particle state and a spinless core, as in |10Be(0+)⊗ν1s1/2〉. Equation (10) can then be expressed in a simpler form as

Equation (11)

where ℓi,ℓf are the orbital angular momenta of the valence neutron in the initial state, and the scattered neutron in the final state, respectively [4951]. ϕi is the radial wave function involved in Φi, and ϕf is the radial component of the final scattering state of the neutron. The E1 effective charge eE1eff for the one-neutron halo nucleus is Ze/A. Although the analysis is to be performed with a distorted wave for ϕf due to the FSI between the core and the neutron, it is instructive to consider the plane wave approximation

Equation (12)

where jf(qr) is the spherical Bessel function. This implies that the above equations (equations (10) and (11)) have the form of the Fourier transform of rϕi(r). Hence, the B(E1) distribution at low q (low Erel) is an amplified image of the density profile of the valence neutron, i.e. the halo state. The amplification is attributed to the factor 'r' in the E1 operator, as well as to the phase space factor 4πr2 in the integrand. The theoretical calculation in [49] shows that B(E1) at low Erel is almost solely determined by the density distribution outside of the range of the mean-field potential.

This interpretation in terms of a Fourier transform also implies that the spatial decoupling of the halo and the core is important. The spatially decoupled halo state overlaps very well with the E1 operator which further emphasizes large separations of the valence neutron and the core. Such an intermediate state has a large overlap with the final scattering state. It is then natural that for stable nuclei, which have no such spatial decoupling, only a high-energy E1 photon can separate the whole neutron fluid from the proton fluid which is realized as GDR.

3.1.3. Spectroscopic significance of soft E1 excitation of one-neutron halo nuclei.

The direct breakup of the one-neutron halo nucleus has led to the idea that the Coulomb breakup can now be used as a spectroscopic tool. For 11Be, Φi is represented by equation (7), and the strong sensitivity of the low-energy E1 to the tail part of the radial wave function (the first term of equation (7)) restricts the contribution to the soft E1 excitation only to the first term (halo term) of equation (7). This allows us to extract the spectroscopic factor α2. The spectroscopic factor extracted in [1315] was in the range α2 = 0.6–0.8. For instance, Fukuda et al [15] extracted the value to be 0.72(4), which is consistent with the values from the transfer reactions [35, 36, 38] and the knockout reaction [37].

Not only the amplitude, but also the spectral shape of B(E1) is useful, as demonstrated below for 19C. Table 1 shows the dependence of the spectral shape on Sn and the angular orbital momentum ℓ in the plane wave approximation [5052]. These two parameters are important since the halo is formed as a result of the quantum tunneling. The ℓ value determines the height of the centrifugal barrier, and it is known that only s and p neutrons can form the halo states [53]. Note that the spectral shape is dependent also on the angular momentum of the final state, ℓf. Another important feature is that the peak is very sensitive to Sn. The lower the Sn value, the lower the energy the peak shifts. This is nothing but the revelation of the soft E1 excitation.

Table 1. Characteristic features of the spectral shapes of dB(E1)/dErel depending on Sn, the orbital angular momentum of the valence neutron (ℓ) in the initial state, and that of the scattered neutron (ℓf) in the final state. The calculation is performed in the plane wave approximation [5052].

ℓ→ℓf dB(E1)/dErel (Erel∼0 MeV) Peak of dB(E1)/dErel
  Ef+1/2rel  
s→p E3/2rel $E_{\mathrm { rel}}=\frac {3}{5}S_{\mathrm {n}}$
p→s E1/2rel Erel≃0.18Sn
p→d E5/2rel $E_{\mathrm { rel}}=\frac {5}{3}S_{\mathrm {n}}$
d→p E3/2rel $E_{\mathrm { rel}}=\frac {5}{3}S_{\mathrm {n}}$

3.1.4. E1 versus E2 excitation for halo nuclei.

As shown in figure 6, the photon numbers vary according to the excitation energy, incident energy and the multipolarity of the excitation. The E1 dominance of the Coulomb breakup of the neutron-drip-line nuclei can be attributed not only to this photon number but also to the effective charge involved in the matrix element. For the direct breakup of a one-neutron halo nucleus, the matrix element is shown in equation (10), which involves the effective charge Ze/A. In general, the same effective charge should be involved in any transition matrix, when the E1 transition is associated with the halo state since the transition occurs for the halo neutron relative to the c.m. of the core. In other words, the cluster-like structure of the nucleus is the cause of the electromagnetic transition to separate the halo part from the core part. Hence, the cross section is roughly proportional to the square of the effective charge.

The effective charge for the Eλ transition for a projectile with a cluster-like structure AZ = (A1Z1 + A2Z2) is written as involved in equation (3.4.19) of [33],

Equation (13)

where the βi corresponds to the parameters describing the position and orientation of the cluster i from the center of mass of the projectile, i.e. $\vec {r_1}= \beta _1 \vec {r}$ and $\vec {r_2}= -\beta _2 \vec {r}$ , where $\vec {r}$ is the relative vector directing from 2 to 1. Here, β1 = A2/A and β2 = A1/A.

For the one-neutron halo nucleus, A1 = 1, Z1 = 0, A2 = A − 1 and Z2 = Z, then, the E1 and E2 effective charges squared are, respectively, written as

Equation (14)

Equation (15)

The formulae explain why the E1 transition is dominant over the E2 transition for neutron halo nuclei. For instance, for 11Be, the E1 effective charge squared is larger by two orders of magnitude (∼A2) than that of E2. The dominance of E1 transition is confirmed for Coulomb breakup of 11Be [13] and 19C [18] by the angular correlation of the fragment in the c.m. of the fragment + neutron. The one obtained for 11Be is shown in figure 11, where the squared sine function expected for the E1 transition is clearly seen. The theoretical curve shown is based on the first-order perturbation theory of the Coulomb breakup [54, 55].

Figure 11.

Figure 11. Azimuthal angular distribution of 10Be in the 11Be rest frame. The solid curve is the result for the pure E1 transition based on the first-order perturbation theory [54, 55].

Standard image

On the other hand, for the proton halo (proton cluster) and α-cluster nucleus, E2 transitions are comparatively important. For instance, in the case of 16O + γ → 12C + α, which is one of the key stellar reactions, the eE1eff = 0, so that the E2 transition becomes dominant. For the proton + core system, such as the one-proton halo nucleus, the effective charges for E1 and E2 transitions are

Equation (16)

Equation (17)

For 8B + γ → 7Be + p, the key reaction of the pp chain in the sun and thus important in estimating the solar neutrino flux, the effective charge squared of E2 (0.68 e2) is larger than that of E1(0.14e2). Due to the photon number spectrum for the low excitation energies, the E2 transition in this reaction is still not significant compared to the E1, but not negligible anymore. The point is that the proton-halo-like case needs to take into consideration the E2 mixture with the E1, but for neutron-halo nuclei the E1 transition is by far dominant, so the E2 transitions can thus be neglected.

3.1.5. Application of Coulomb breakup to the structure study—soft E1 excitation of 19C.

Here we show two applications of Coulomb breakup, which are on 19C and 15C. In this subsection, we discuss the case of 19C. Before the Coulomb breakup experiment [18], the structure of 19C had been controversial partially because of a large uncertainty in the direct mass measurements ranging from almost 0 to 700 keV [5659]. The mass evaluation in 1993 based on these direct mass measurements was 0.16 ± 0.11 MeV [60]. Such a small Sn value for 19C suggested a possible halo state. However, one should note that the condition to form a neutron halo is not only the smallness of the Sn value, but also the smallness of the orbital angular momentum of the least bound neutron. For 19C, it was not clear if its valence neutron satisfies that condition. Since the 13th neutron in 19C occupies an orbital in the sd shell, the following two possibilities for the ground state configuration should be considered:

Equation (18)

Equation (19)

where α, β, γ and δ denote the spectroscopic amplitude for each configuration. It is interesting to note that the shell model calculations [61, 62] suggest that these two states are almost degenerate in energy.

A kinematically complete measurement of the Coulomb breakup of 19C at 67 MeV per nucleon was carried out at RIKEN [18]. Figure 12 (left) shows the Coulomb breakup cross section as a function of Erel. Here, the nuclear breakup component was subtracted using the breakup spectra for a C target, assuming that the breakup cross section with the light target arises from the nuclear breakup origin, and the nuclear breakup cross section is scaled by the target + projectile radii. The method of selecting the forward angles, used in more recent 11Be data [15] shown above was not adopted in these data due to lower statistics. The Coulomb breakup spectrum in figure 12 (left) shows an asymmetric shape with a large cross section of 1.19 ± 0.11 barn (b), whose large amplitude and the shape clearly show the halo characteristics.

Figure 12.

Figure 12. Left: Coulomb breakup cross section spectrum obtained for 19C at 67 MeV per nucleon on a Pb target at RIKEN [18]. The dot-dashed and dashed curves are calculations for the configurations, |18C(2+)⊗ν1s1/2〉 and |18C(0+)⊗ν0d5/2〉, respectively, of Jπ = 5/2+ assignment with unit spectroscopic factors and Sn = 160 keV. The dotted and solid curves are calculations for the configuration |18C(0+)⊗ν1s1/2〉 with Sn = 160 and 530 keV, respectively. Right: angular distribution of the Coulomb breakup cross section of 19C on a Pb target. Figure reprinted with permission from Nakamura et al [18]. Copyright (1999) by the American Physical Society.

Standard image

The data are compared with the direct-breakup calculations for the possible four configurations, as shown in figure 12 (left). It is shown that good agreement is obtained for a higher Sn value of 530 keV for |18C(0+)⊗ν1s1/2〉 with Jπ = 1/2+. This Sn value 530 keV adopted here was obtained from an independent analysis using the angular distribution, as shown below. The large spectroscopic factor of α2 = 0.67 was also deduced. The obtained configuration matches the prediction by a shell model calculation with the WBP interaction [61, 63]. The dominance of a weakly bound s-wave valence neutron shows that 19C is another well-established case of the 1n halo nucleus as in 11Be.

The mass of 19C (more precisely, Sn) was extracted from the Coulomb breakup for the first time in this experiment [18]. The result has been indeed incorporated into the newer mass evaluation [64]. Here, a feature of Coulomb breakup, namely that the angular distribution of 19C (18C + n system) is sensitive to Sn, was used. As in equation (8), the angular distribution of the cross section is proportional to that of the E1 virtual photon (dNE1/dΩ). The point is that the angular distribution of the photon is a function of Ex, which is related to Sn via Ex = Erel + Sn. Note that the determination of Sn from the angular distribution is less model dependent compared to the determination of Sn from the peak position of the B(E1) spectrum which needs to assume the spectral shape to be the pure direct breakup.

Figure 12 (right) shows the measured angular distribution in comparison with the calculated spectra, where an angular resolution of 8.4 mrad (in 1σ) is folded. The value Sn = 530 ± 130 keV gives a best fit to the distribution, whereas the previously adopted value of Sn = 160 keV did not reproduce the angular distribution. This higher Sn value was confirmed by the measurement of the momentum distribution of 18C in coincidence with the de-excitation γ-ray from the first excited state of 18C at MSU [65].

3.1.6. Application of the Coulomb breakup to the radiative capture reaction 14C(n,γ)15C of astrophysical interest.

Coulomb breakup is also useful in extracting the radiative neutron capture reaction (n,γ), which is the inverse of the (γ,n) reaction. Coulomb breakup has been used in the application to nuclear astrophysics [66, 67], not only for the cases relevant to the neutron halo, but also and more widely for the proton-rich nuclei, i.e. proton radiative capture.

Here, we show the case of the Coulomb breakup of 15C whose inverse process is 14C(n,γ)15C.

The following equation based on the principle of the detailed balance relates the photo-absorption cross section (σγn) to the radiative neutron capture cross section (σnγ), i.e.

Equation (20)

where IA = 1/2 and IA−1 = 0 for the present case, and μ denotes the reduced mass of 14C + n, and Ec.m. is the c.m. energy of n + 14C which is equivalent to Erel.

Due to the phase-space factor, the photo-absorption cross section is significantly larger than the inverse process. For the current case with Ec.m. = 0.5 MeV, this ratio is very large, as σγn(Eγ)/σnγ(Ec.m.) ≃ 150. The Coulomb breakup has an additional advantage that the photon numbers are further multiplied, which is of 2–3 orders of magnitude. Furthermore, the kinematic focusing and the availability of a thick target of the order of 100 mg cm−2–1 g cm−2, are also advantages, which can compensate for a weak intensity of the RI beam. Note also that the Coulomb breakup of 15C can avoid the difficulty of using the radioactive 14C target as well.

The neutron capture reaction on 14C is considered to be important in the nucleosynthesis possibilities, such as the neutron-induced CNO cycle [68], revised r-process scenario initiating from the lightest elements [69, 70] and the inhomogeneous big-bang nucleosynthesis [71]. The Coulomb breakup of 15C is also important in terms of the spectroscopic significance of the 15C nucleus itself. 15C has a moderate-sized neutron halo with a neutron separation energy Sn of only 1.218 MeV. Its main configuration is |14C(0+)⊗ν1s1/2〉 upon the N = 8 core of 14C. Coulomb breakup of 15C is expected to proceed via soft E1 excitation due to the direct breakup mechanism.

In such a situation, the direct neutron capture process for a p-wave neutron [50] is considered to be a dominant process in the inverse reaction. This particular capture process is called direct neutron capture since the neutron in the continuum is captured onto the 14C nucleus, as the inverse process of direct Coulomb breakup. The p-wave capture process, as a stellar reaction, is an exceptional case, considering that the neutron capture of a low-energy s-wave neutron which follows the 1/v law is usually dominant.

Another important feature of this Coulomb breakup is that the inverse reaction of the neutron capture of 14C has been measured [72, 73], and one can thus evaluate the validity of the Coulomb breakup (or vice versa). There were three Coulomb breakup measurements [16, 17, 74]. The latter two [16, 17] are consistent with the newer neutron capture measurement [73]. Here we introduce the most recent Coulomb breakup result [17], where the experiment was done at 68 MeV per nucleon at RIKEN.

Figure 13 (left) shows the Coulomb breakup cross sections of 14C + n as a function of Erel obtained in the breakup of 15C on a Pb target at 68 MeV per nucleon. For extracting the Coulomb breakup component, we adopt the same procedure as in the Coulomb breakup of 11Be [15], using the angular selection at forward angles. The solid squares show the breakup cross section integrated over the scattering angular range 0° ⩽ θ ⩽ 6.0° which nearly corresponds to the whole acceptance. Open circles show the cross section for a selected angular range, 0° ⩽ θ ⩽ 2.1°, which corresponds to the impact parameter range b > 20 fm.

Figure 13.

Figure 13. Left: relative energy spectra for Coulomb breakup of 15C. Solid squares represent the data for scattering angles up to 6°, and open circles represent the data for selected scattering angles up to 2.1°. The solid curves are calculations for a direct breakup model with a spectroscopic factor α2 = 0.91, for the halo configuration (a = 0.5 fm and r0 = 1.223 fm). Right: B(E1) spectrum for 15C excitation. The solid curve corresponds to the direct breakup model calculation. The dot-dashed curve is the same calculation, before folding with the experimental resolution. Figure reprinted with permission from Nakamura et al [17]. Copyright (2009) by the American Physical Society.

Standard image

The breakup cross sections integrated up to Erel = 4 MeV are 670 ± 14(stat.) ± 40(syst.) mb and 294 ± 12(stat.) ± 18(syst.) mb for the whole acceptance (0–6°) and the selected angular range (0–2.1°), respectively. Although the spectral shape is attributed to Coulomb breakup of halo nuclei, the cross section is about 1/5–1/3 of those for the conventional halo nuclei such as 11Be [15], and 19C [18]. This indicates that the size of the halo in 15C is not as large as in the more weakly bound halo nuclei. The B(E1) spectrum has been extracted for the angular selected data, and is shown in figure 13 (right), showing excellent agreement with the direct breakup calculation, as shown in the solid curve. The dot-dashed curve shows the calculation before folding the experimental resolutions.

Figure 14 shows the neutron capture cross section σnγ extracted from the B(E1) spectrum by applying the principle of detailed balance (equation (20)). Excellent agreement is obtained with the p-wave direct radiative capture model calculation [50]. This result is consistent with the consideration that the final state of the 14C capture reaction (namely 15C g.s.) is a halo state with a dominant s-wave component. The result is also found to be consistent with the neutron capture measurement in [73]. The overall agreement of the Coulomb breakup result with the direct capture measurement suggests that Coulomb breakup can be a good alternative for obtaining the neutron capture cross sections involving radioactive nuclei. One should note, however, that in the inverse reaction of Coulomb breakup, one can restrict the reaction to the ground state. For 15C, there is only one excited state at 0.74 MeV, and the neutron capture to this state is estimated to be negligible [46, 71, 75]. Although such fortunate situations may be scarce for heavier neutron-rich nuclei, we may expect lower level densities near closed shells, such as N = 50 and 82, where Coulomb breakup can be applicable. These nuclei may be relevant to the r-process, and thus may also be of importance.

Figure 14.

Figure 14. Neutron capture cross section of 14C which leads to the 15C ground state. Solid black circles represent the results of the current experiment, and the red open squares represent those from the direct capture measurement [73]. The dot-dashed curve is the calculation based on the direct radiative capture model, whereas the solid curve is the same calculation but includes experimental resolution. Both calculations were performed using a spectroscopic factor α2 = 0.91 and with potential parameters a = 0.5 fm and r0 = 1.223 fm. Figure reprinted with permission from Nakamura et al [17]. Copyright (2009) by the American Physical Society.

Standard image

3.1.7. Inclusive Coulomb breakup of 31Ne.

The experimental efforts of Coulomb breakup on halo nuclei are now directing toward the heavier nuclei. As shown in the nuclear chart in figure 7, nuclear species investigated by the Coulomb breakup have been limited to the very light neutron-rich nuclei as far as the halo physics is concerned. However, the halo could be formed in the heavier nuclei. The questions are how the halo structure appear in heavier neutron-drip-line nuclei and whether halo states can appear in many regions of the nuclear chart or not. These are not trivial since in heavier regions configuration mixing is more complex, higher angular momentum is involved in the structure and a variety of deformed shapes is expected. Here, we show the first experiment of Coulomb breakup of nuclei in the heavier region over Z = 8 or over the sd neutron shell. Before this experiment, the Coulomb breakup has been measured only up to 19C, and has for long been the heaviest halo nucleus experimentally established.

The beam intensity of the neutron-drip-line nuclei in this region has been limited. One breakthrough was the start of the new-generation RI beam facility RIBF at RIKEN, whose layout is shown in figure 15. The final goal of this facility is to obtain all the heavy ion beams up to 238U at the intensity over 1 pμA (particle micro-ampere), corresponding to ∼6 × 1012 particles per second. So far, for instance, a rather high maximum intensity of 250 pnA has already been achieved for 48Ca. With this intense 48C beam, one can now have access to neutron-rich nuclei nearly up to the neutron-drip line for B–Si.

Figure 15.

Figure 15. The layout of the RIBF at RIKEN as of 2012, which represents the world's largest RI beam facility in the world at this moment. The sequence of the accelerators, for instance, RILAC, RRC, FRC, IRC and SRC, is used for the acceleration up to 345 MeV per nucleon.

Standard image

As demonstrated above, Coulomb breakup reaction is a powerful tool for investigating 1n halo nuclei in mapping the dB(E1)/dErel function by a kinematically complete measurement. However, as a first step, the inclusive Coulomb breakup may also be important, due to the simplicity and the feasibility with a smaller beam intensity. For 31Ne, we measured the inclusive Coulomb breakup cross section at 230 MeV per nucleon [34] at RIBF at RIKEN. The beam intensity was only about 5 counts per second with the 48Ca beam at 345 MeV per nucleon with typical 60 pnA at that time.

Since the first observation of 31Ne in the late 1990s [76], 31Ne has been the next heavier candidate for a 1n halo nucleus due to the small 1n separation energy. According to the mass evaluation in 2003 [64], Sn was theoretically estimated as 0.332 ± 1.069 MeV, while the recent direct mass measurement showed Sn to be 0.29 ± 1.64 MeV [77]. Experimental studies of 31Ne are thus important as to how and in what form heavier halo nuclei appear. Once the halo structure of 31Ne is established, we may obtain a key to understand how a single-particle configuration plays a role in halo formation. 31Ne (N = 21) is also interesting in that it may be inside the island of inversion where shell model calculations showed that N = 20 magicity is lost and significant 2p–2h configurations are mixed [7880]. Note that the neighboring nuclei 32Mg [81], 32Na [82] and 30,32Ne [8384] were found within this island experimentally.

It should be noted that for the conventional shell order, where the valence neutron resides in the 1f7/2 orbital upon the 30Ne core, a halo never develops since the tunneling effect is blocked by the high centrifugal barrier. The halo formation of 31Ne is only possible when a strong shell modification occurs, such that the 1p3/2 orbital lowers below the 0f7/2 orbital. As mentioned in section 2.2, the halo signature can be obtained by the inclusive cross section. For 31Ne, the significant cross section of about 0.5 b or more can be obtained only when the soft E1 excitation occurs.

Figure 16 (left) shows the 1n removal cross sections of 19,20C and 31Ne on Pb and C targets obtained in this experiment. We readily find that the cross sections of 19C and 31Ne are both significantly larger than that of 20C. Secondly, the ratio of the cross section for Pb to that for C is 9.0 ± 1.1 for 31Ne and 7.4 ± 0.4 for 19C, much larger than the ratio for nuclear breakup only, which is estimated to be about 1.7–2.6. This demonstrates that the cross section for a Pb target is dominated by Coulomb breakup for 19C and 31Ne.

Figure 16.

Figure 16. Left: one neutron removal cross sections for 19C, 20C and 31Ne on Pb (diamonds) and C (circles), and extracted Coulomb breakup cross sections (squares) obtained at about 230 MeV per nucleon. The large Coulomb cross section of 31Ne close to that of 19C (established halo nucleus) indicates the 1n halo structure of 31Ne. The Coulomb breakup cross section for 19C estimated from the kinematically complete measurement [18] is shown by the green square. The data for 19C and 31Ne are taken from [34]. The data for 20C are preliminary. Right: the Coulomb breakup cross section for 31Ne on Pb at 234 MeV per nucleon is compared with direct-breakup calculations for the configurations shown with C2S = 1. An example of lower C2S value (C2S = 0.5) is also shown by the dot-dashed curve for the 1p3/2 configuration.

Standard image

The Coulomb breakup component of the 1n removal cross section on Pb was deduced by subtracting the nuclear component estimated from σ−1n(C). The Coulomb breakup cross section for 31Ne was thus obtained to be σ−1n(E1) = 540 ± 70 mb. The dominance of the Coulomb breakup for the reaction of 31Ne on Pb and the deduced σ−1n(E1) of some 0.5 b, nearly as high as the established halo nucleus 19C, indicates the occurrence of soft E1 excitation for 31Ne, which provides evidence for 1n halo structure of 31Ne.

According to the direct breakup mechanism, the single-particle structure of the ground state of 31Ne can be examined. Figure 16 (right) compares the experimentally deduced σ−1n(E1) with calculations for possible valence–neutron configurations as a function of Sn. Since there is a large experimental uncertainty in the Sn value of 31Ne, the calculations were shown as a function of Sn. The results of the direct-breakup calculations for the valence neutron either in the 1s1/2, 0d3/2, 0f7/2 or 1p3/2 orbital, being coupled to the ground state of 30Ne with C2S = 1. A more detailed analysis including possible configurations of the valence neutron coupled to the 30Ne(2+1) state is presented in [34]. The comparison in figure 16 (right) shows that the data are reproduced by the configuration of |30Ne(0+1)⊗ν1p3/2〉 (Jπ = 3/2) or |30Ne(0+1)⊗ν1s1/2〉 (Jπ = 1/2+), while not by |30Ne(0+1)⊗ν0d3/2〉 or |30Ne(0+1)⊗ν0f7/2〉. The significant contribution of such a low-ℓ valence neutron is again consistent with the 1n halo structure in 31Ne. The other important implication of this result is that the conventional shell model configuration of |30Ne(0+1)⊗ν0f7/2〉 for the N = 21 nucleus does not represent a primary configuration of the 31Ne ground state. Large-scale Monte-Carlo shell model (MCSM) calculations employing the SDPF-M effective interactions [80] support the assignment of Jπ = 3/2 having a |30Ne(0+1)⊗ν1p3/2〉 contribution, which is consistent with the current findings.

The shell model calculations (MCSM) also showed the large configuration mixing, where |30Ne(0+1)⊗ν1p3/2〉 could be mixed with |30Ne(2+1)⊗ν1p3/2〉 and |30Ne(2+1)⊗ν0f7/2〉. Such a large configuration mixing may be described in terms of deformation. Recently, the current data were interpreted by a deformed mean-field model [85]. There, the 21st neutron (1p3/2) can be orbiting in a deformed mean-field potential, namely in the Nilsson levels [330]1/2 or [321]3/2, although the possibility of an s-wave valence neutron ([200]1/2+) in a strongly deformed mean field (β > 0.59) was not fully excluded. More recently, the ground state properties of 31Ne were also discussed using the particle-rotor model, which takes into account the rotational excitation of the 30Ne core [86]. It is interesting to note that halo properties, such as direct breakup dynamics and soft E1 excitations, are kept for the wave functions obtained in a deformed mean-field potential.

One neutron halo structure of 31Ne has recently been confirmed also by the reaction cross section measurement of this nucleus at RIBF [87]. Further experimental work on determining the content of the halo structure of 31Ne is highly called for. In particular, it is important to determine Sn as well as the configuration mixing experimentally. For that, a kinematically complete measurement of 31Ne would be one of the key experiments, where C2S and Sn can be extracted unambiguously for this 1n halo nucleus. At RIBF at RIKEN, such experiments are expected to be performed using the SAMURAI (Superconducting Analyser for MUlti-particles from Radio-Isotope beam) facility (see figure 20 in section 3.3).

3.2. Soft E1 excitation of 2n halo nuclei

The soft E1 excitation of 2n halo nuclei has a different aspect from that of 1n halo nuclei, i.e. it has two neutrons involved in the initial state and excitation. The three-body nature makes the interpretation of the experiment difficult. As will be mentioned, the two neutron correlation in the initial state and the FSIs between n–core and n–n will play important roles.

The first Coulomb breakup measurement for halo nuclei was made for 11Li in an inclusive manner, which showed significant enhancement of low-energy dipole strength as mentioned in [5]. Later on, final-state exclusive, kinematically complete measurements of Coulomb breakup were made for 6He [19, 88], 11Li [20, 3941] and 14Be [89]. Here, the kinematical complete measurements of 6He and 11Li are primarily described in order to clarify the characteristic features.

3.2.1. Coulomb breakup and soft E1 excitation of 11Li.

In the 1990s, kinematical complete measurements for 11Li were made at 28 MeV per nucleon at MSU [39], at 43 MeV per nucleon at RIKEN [40] and at 280 MeV per nucleon at GSI [41] (see figure 17 (right)). However, these results are controversial with each other. More recently, new data on 11Li were obtained [20] with much higher statistics and with higher sensitivity for two-neutron detections.

Figure 17.

Figure 17. Left: the experimental setup for the Coulomb breakup measurement of 11Li on Pb at 70 MeV per nucleon at RIKEN. The setup contains a dipole magnet (MAG), drift chamber (FDC), hodoscope (HOD) and two walls of neutron detector arrays (NEUT-A,B). Right: B(E1) spectrum obtained in the Coulomb breakup of 11Li at 70 MeV per nucleon [20] compared with the previous data obtained at MSU at 28 MeV per nucleon (dot-dashed-line) [39], RIKEN at 45 MeV per nucleon (solid histogram) [40] and at GSI at 280 MeV per nucleon (zone between dashed lines) [41]. The data are also compared with the three-body calculation shown by the solid curve [45]. Figures reprinted with permission from Nakamura et al [20]. Copyright (2006) by the American Physical Society.

Standard image

The controversial results on 11Li are considered to arise from the difficulty of the multi-neutron measurements. The fast-neutron measurement made above uses plastic or liquid scintillators, where the recoiled protons or knocked-out protons in the detector ingredients are a major source of the detection. In this process, this neutron that induced this recoil particle is scattered, and if it induces another signal at a different part of the detector, then the hit signals detected become double. Such an event is called cross talk, which mimics the true two-neutron hit.

The Coulomb breakup measurement of 11Li on Pb at 70 MeV per nucleon at RIKEN [20] used the setup shown in figure 17 (left), which took care of the cross talk events as follows. The neutron detectors composed of 54 rods of plastic scintillators (214(H× 6.1(V× 6.1(D) cm3 each) were arranged into two walls (12 × 2 rods for the front (NEUT-A) and 15 × 2 rods for the rear (NEUT-B)), separated by 1.09 m. This separation is essential to distinguish the two neutrons by imposing the cross-talk-cut condition. When two hits are detected in NEUT-A and NEUT-B independently, one can reconstruct the velocity vA for the hit in A, and the apparent velocity vAB between the two walls from the hit timing of each wall. Then we discard all the events with vA ⩾ vAB, thereby excluding cross talk events since the scattered neutron has less velocity than that of the incoming neutron. This measurement also uses the two hits in the same wall (see [20, 47] for details). This revised measurement enhanced significantly the sensitivity down to Erel ∼ 0 MeV in the B(E1) spectrum.

Figure 17 (right) shows the B(E1) distribution in this experiment, which is compared with the previous three data sets. The result shows substantial E1 strength at very low Erel around 0.3 MeV. This feature is in contrast to the previous data, which showed much reduced strength toward low relative energies. This is considered due to the lower neutron detection efficiency near Erel ∼ 0 MeV in those experiments.

The spectrum amounts to an energy-integrated B(E1) strength of 1.42 ± 0.18 e2 fm2 (4.5(6) W.u.) for Erel ⩽ 3 MeV, which is the largest soft E1 strength ever observed. In fact, this E1 strengths is about 40% larger than that for the one-neutron halo nucleus 11Be [14, 15]. The extraction using the equivalent photon method depends on the S2n value since the virtual photon spectrum depends on Ex(=Erel + S2n). Nakamura et al [20] adopted S2n = 0.3 MeV from the 2003 mass evaluation [64]. When one uses the most recent precise mass measurement result of 11Li (S2n = 0.36915(65) MeV) [90] using the Penning trap technique, the E1 strength for 11Li is reevaluated as 1.49(19) (e2 fm2) for Erel ⩽ 3 MeV, or about 5% larger. Such a slight change within the experimental uncertainty does not change the conclusion, however.

Figure 17 (right) also compares the data with the three-body model calculation of Esbensen and Bertsch [45], which includes the two-neutron correlations in both the initial and final states in the E1 excitation. This calculation reproduces the data very well with no adjustment of normalization. Recently, a revised three-body calculation by Esbensen et al [91], which took into account the recoil effect, also showed overall agreement with the data, although some deviation occurs around Erel ∼ 0.5 MeV (see figure 2 of [91]). This deviation may indicate the effect of the core excitation or other effects. The recent calculation by Myo et al [92], which incorporated the effect of the core polarization using the tensor optimized shell model, reproduced better the peak region. These overall agreements of both the spectral shape and absolute strength in these calculations indicate the presence of a strong two-neutron correlation in 11Li, in both the ground and final states (FSIs).

3.2.2. Coulomb breakup and soft E1 excitation of 6He.

6He exhibits a Borromean structure composed of α + n + n. The S2n value is 972.44(66) keV [64], larger than that of 11Li by a factor of about 3. The shell configuration of the two halo neutrons in 6He is p dominant, while that of 11Li has a mixture of s and p. The other difference is that the core of 6He is a rigid 4He with Jπ = 0+, while that of 11Li is a soft 9Li with an odd spin (Jπ = 3/2). Hence, it is expected that the 6He Coulomb breakup could provide a more rigorous test of the three-body- and reaction models of the soft E1 excitation of 2n halo nuclei.

For the Coulomb breakup of 6He, two kinematically complete measurements at 25 MeV per nucleon on U at MSU [88] and at 240 MeV per nucleon on Pb at GSI [19] were made. The B(E1) spectra from these two data show apparent different behavior in strength and in shape. This difference could be due to the different sensitivity of the two data sets. Since the experiment at MSU [88] had a much smaller acceptance which cut the spectra at higher excitation energies, we hereafter discuss the experiment at GSI [19].

The result of GSI is shown in figure 18 for the cross section and the extracted B(E1) spectrum using the equivalent photon method. The integrated B(E1) sum is 0.59(12) e2 fm2 (2.8(6) W.u.) for Ex ⩽ 5 MeV and 1.2(2) e2 fm2 (5.7(10) W.u.) for Ex ⩽ 10 MeV. Compared to 11Li, the absolute value of E1 strength is smaller, and the spectrum shows a broader peak, extending over 6 MeV in Erel.

Figure 18.

Figure 18. Left (top): breakup cross sections of 6He for Pb and C targets obtained at 240 MeV per nucleon [19]. Left (bottom): corresponding correlation functions (for details, see [19]). Right: extracted B(E1) spectrum for 6He, which is compared to the theories of [93] (dotted curve) and of [94] (dashed curve). Figures reprinted with permission from Aumann et al [19]. Copyright (1999) by the American Physical Society.

Standard image

Qualitatively, such a difference may be attributed to the fact that 6He is more bound than 11Li, and is p-wave dominant which has a finite centrifugal barrier even though it is small. Theoretically, there are quite a few theoretical publications for the electromagnetic excitation of 6He [93103]. Theoretical interpretations of the data are still not fully established, and in this review we cannot fully trace those aspects. We introduce here one recent theoretical study by Kikuchi et al [98], which investigated separately the effects of direct breakup into non-interacting three-body particles and those of initial and FSIs.

Figure 19 shows the comparison in [98] with the data, where the left figure shows the comparison with the cross section dσ/dEx, whereas the right two figures show the cross sections in terms of the two-body relative energies α + n and n + n. We find overall agreement for the three-body decay spectrum, although a slight enhancement around Erel ∼ 1 MeV and hindrance over 2 MeV in the theoretical curve is seen. On the other hand, better agreement is achieved for the two-body decay spectra. In particular, the α + n data are in excellent agreement showing a strong effect of the 5He(3/2) ground state, i.e. the effect of the FSI between α and neutron. Such an effect was already pointed out by the data analysis in [19]. Kikuchi et al [98] also found the effect of the strong nn FSI, while the nn correlation in the initial ground state affects negligibly the spectrum. In the next section, we discuss the non-energy weighted cluster sum rule, where the ground state nn correlation can be revealed in the summed E1 strength.

Figure 19.

Figure 19. Left: Coulomb breakup cross section of 6He at 240 MeV per nucleon [19] is compared with the three-body calculation of Kikuchi [98]. Middle (a): the two-body 4He + n relative energy spectrum in the Coulomb breakup of 6He is compared with this theory. Right (b): the two-body n–n relative energy spectrum in the same reaction and its comparison with this theory. Figures reprinted with permission from Kikuchi et al [98]. Copyright (2010) by the American Physical Society.

Standard image

3.2.3. Non-energy weighted cluster sum rule for 2n halo nuclei and dineutron correlation.

The spatial two-neutron correlation in the ground state (initial state) of two-neutron halo nuclei can be quantitatively estimated by the non-energy weighted E1 cluster sum rule [45, 104]. Such a sum rule was discussed for 1n halo nuclei as in equation (9). For the case of a two-neutron halo nucleus, the E1 operator $\vec {r}$ is replaced by $\vec {r_1}+\vec {r_2}$ , where $\vec {r_1}$ and $\vec {r_2}$ are the position vectors of the two valence neutrons relative to the c.m. of the core. Namely, the sum rule is written as

Equation (21)

Here, $\vec {r_1}+\vec {r_2}$ can be related to the rms of rc,2n, which is the distance from the center of the core to the c.m. of the two halo neutrons. Note that the term of $\vec {r_1}\cdot \vec {r_2}$ involves the opening angle θ12 between the two position vectors of the two valence neutrons, i.e. the information on the two-neutron correlation. The value of 〈rc,2n〉, and hence B(E1), becomes larger for smaller spatial separation of the two neutrons when θ12 approaches 0°. The integrated B(E1) thus provides a good measure of the two-neutron spatial correlation.

For the 11Li data in [20], the integral of the E1-response calculation up to Erel = 3  MeV accounts for about 80% of the total E1 cluster sum-rule strength when one assumes the extrapolation according to the calculation in [45]. The estimated B(E1) strength of 1.78 ± 0.22 e2 fm2 for Ex ⩾ S2n is thus deduced, which corresponds to $\sqrt {\langle r^2_{c,2\mathrm {n}}\rangle } = 5.01 \pm 0.32\,{\mathrm { fm}}$ , which is consistent with other theoretical estimations of 5–6 fm [91, 92, 107]. The charge radius measurement of 11Li extracted $\sqrt {\langle r^2_{c,2\mathrm {n}}\rangle }=5.97(22)$  [91, 105, 106]. The slight deviation between the value from the Coulomb breakup and the charge radius could be due to a possible core polarization effect [91].

Assuming the sum rule value of 1.07 e2 fm2 for the two non-correlated neutrons as calculated in [45], the total experimental strength of 1.78 ± 0.22 e2 fm2 can be translated to 〈θ12〉 = 48+14−18 degrees. This angle is significantly smaller than the mean opening angle of 90° expected for the two non-correlated neutrons. Hence, an appreciable two-neutron spatial correlation is suggested for the two halo neutrons. An important feature of the soft E1 excitation of 2n halo nucleus is that the larger polarization of the charge due to a stronger dineutron-like correlation enhances the soft E1 excitation.

It can be shown that the spatial nn correlation is related to a mixture of different parity states in the valence neutrons of 11Li. In a simple shell model picture where the inert 9Li core is assumed, the 11Li could be described as

Equation (22)

Here, we are interested in the mean opening angle between $\vec {r_1}$ and $\vec {r_2}$ . The expectation value of cos θ12 can be written as

Equation (23)

Equation (24)

The terms for α2 and β2 are null since these are odd functions of cos θ12. Hence, a non-zero expectation value implies a mixture of these different parity configurations. Namely, no mixture of different parity states (either α2 = 0 or β2 = 0) gives rise to 〈θ12〉 = 90° and no 2n correlation, while the mixture leads to the 2n spatial correlation.

A similar discussion can be applied also to 6He. The authors of [19] showed that the B(E1) strength up to Ex = 10 MeV in 6He is translated into $\sqrt {\langle r^2_{c,2\mathrm {n}}\rangle }=3.36(39)\,{\mathrm { fm}}$ . The opening angle estimated in [107] using this value is 83+20−10 degrees, which is certainly larger than the case of 11Li. The closeness to 90° shows that the two-neutron halo may have a smaller mixture of different parities from p states.

3.3. Perspectives on the soft E1 excitation

We find that the Coulomb breakup of 1n halo nuclei and the revealed enhanced low-energy strength (soft E1 excitation) can be well understood by the direct breakup mechanism, thereby providing a good experimental tool to probe the single-particle halo state. The Coulomb breakup of 2n halo nuclei provides, in addition, information on dineutron correlation for 2n halo nuclei. However, we note that full understanding of Coulomb breakup of 2n halo nuclei is still on the way. There are quite a few theoretical works on the Coulomb breakup on 11Li (see, for instance, [91, 102, 104, 108117]), which are still controversial. For the two-neutron halo case, the non-resonant character seems to persist for 2n halo nuclei except for the strong core-n and nn FSIs. However, for instance, a more recent theory [117] relates the enhanced strength with a strong dipole resonance nature of 11Li. In this sense, the issue whether it is resonance or non-resonance remains to be solved.

Experimentally, it is important to explore other 2n halo cases with different shell properties. The most neutron-rich carbon isotope, 22C, provides such a case, which was recently confirmed as a two-neutron halo nucleus from the reaction cross section measurement [118] and the 2n-removal reaction and the momentum distribution of 20C fragment in the 22C + C reaction at RIBF [119]. The two neutrons in 22C should reside in s and d [118, 119], although the detailed mixture is still controversial and should be investigated further. At RIBF at RIKEN the inclusive Coulomb breakup of 19B and 22C on Pb around 240 MeV per nucleon has recently been measured, and the results show the enhanced cross section of about 1 b, consistent with the revelation of soft E1 excitation. More recently, kinematical complete measurements for 22C have also been made in 2012 at the newly commissioned large acceptance spectrometer SAMURAI as shown in figure 20. We also note that further high-precision measurement for 6He and 11Li is necessary to pin down the nature of Coulomb breakup and soft E1 excitation of 2n halo nuclei. This is also important for obtaining clearer evidence of the dineutron correlation, and its understanding. In the near future, we also note that the 4n halo could be in the scope.

Figure 20.

Figure 20. The schematic view of the SAMURAI facility at RIBF, which was commissioned in 2012. The secondary beam, for instance 22C, comes from the left-hand side of this figure, impinges on the target, located in the inner side of the DALI2 detector array. The breakup particle, for instance 20C, is bent about 60° in the superconducting magnet at about 3 T, detected with the heavy-ion detectors (the two multi-wire drift chambers FDC1 and FDC2, and the plastic scintillator hodoscope HODF). The fast neutrons are detected in NEBULA neutron detectors, which have 120 modules of plastics scintillators, each of which has 12(H× 12(D× 180(V) cm3.

Standard image

4. Giant and pygmy resonances in neutron-rich nuclei

4.1. Experimental investigations on the multipole response of exotic nuclei

The excitation spectra of nuclei are dominated by collective giant resonances of various different character. The question of how the multipole response of nuclei evolves when moving from stable to more and more neutron–proton asymmetric systems has been studied in many theoretical works and also experimentally. For an overview of theoretical studies, see the review article by Paar et al [25]. From the experimental point of view, it is mainly the electric dipole response which has been studied so far due to experimental difficulties. Isoscalar monopole and quadrupole giant resonances, for instance, are studied usually in α scattering experiments at low momentum transfer. In inverse kinematics, this results in very-low-energy recoiling α particles. In order to be able to measure the angular distribution and energy of the recoils, very thin targets, i.e. gas targets, are required implying a low luminosity. An attempt has been made by using an active target at GANIL [120] for an investigation of the GMR and GQR in 56Ni in deuteron scattering. At GSI, a method is being developed to study elastic and inelastic hadron and light-ion scattering in the storage ring by using a gas-jet target [121, 122]. The luminosity that can be reached with this method is much higher and thus will be very well suited for giant-resonance studies in the future.

So far, mainly the process of heavy-ion-induced electromagnetic excitation as discussed above has been employed in the study of radioactive nuclei. In order to excite the nucleus up to energies in the giant-resonance domain, however, rather high beam energies are needed due to the adiabatic cutoff, see above. Typical beam energies used are around 500 MeV per nucleon. The cross section is dominated by electric dipole transitions. Typically, the cross section for exciting the giant quadrupole resonance is about one order of magnitude smaller than the one for exciting the isovector GDR at energies around 500 MeV per nucleon. For high-precision measurements, however, different multipoles have to be considered. Measurements at different beam energies as well as a precise measurement of the scattering-angle distribution can be used to decompose the different multipolarities. The results from the experiments with radioactive beams discussed below have not been corrected for possible quadrupole contributions. In the analysis, however, contributions from E2 excitations have been taken into account by assuming standard parameters for the giant quadrupole resonance.

In addition to possible admixtures of excitations other than E1, nuclear excitations have to be considered. For impact parameters significantly larger than the sum of the two radii of the target and the projectile, only the electromagnetic interaction is important, while for smaller impact parameters, nuclear processes dominate. For the calculation of the electromagnetic cross section, either a minimum impact parameter is considered, or the nuclear processes are explicitly taken into account [123]. Most of the nuclear reaction channels, however, populate final states other than considered here. Only inelastic scattering, i.e. excitation of the projectile to the excitation-energy region of interest with subsequent decay (excluding individual nucleon–nucleon collision with large momentum transfer), contribute to the cross section of interest. In the small impact parameter range, where nuclear inelastic excitation is important, also electromagnetic excitation is competing and the two amplitudes might interfere. Since the integrated cross section for electromagnetic excitation is much larger in the cases considered here, and only a small fraction of the cross sections stems from an impact parameter range where both processes are competing, the effects due to interference are usually small, and in particular for scattering-angle integrated cross sections negligible. In extracting the cross section for electromagnetic excitation, the nuclear contribution is determined usually by a measurement with a carbon target and then subtracted from the total cross section after proper scaling. Typically, the nuclear contribution in the cases discussed here is of the order of 10–20%.

Differential cross sections with respect to excitation energies can be deduced using the invariant-mass method and a measurement of the four momenta of all outgoing particles, i.e. a kinematically complete measurement of the final state. The experimental setup used at GSI is shown schematically in figure 21.

Figure 21.

Figure 21. Schematic view of the LAND experimental setup at GSI. The incoming ion beam is tracked and identified event by event with projectile-tracking detectors. The target is surrounded by a modular NaI detector array (160 crystals) for photon detection and calorimetry (Crystal Ball). Charged fragments are deflected by the large-gap dipole magnet ALADIN and tracked by fragment-tracking detectors which serve in conjunction with energy-loss and time-of-flight measurements (TOF) for fragment identification. Neutrons are detected with an acceptance of typically ±80 mrad around 0° with the LAND detector placed 12 m downstream of the target. From the combined information the excitation energy can be reconstructed by using the invariant-mass method. The inset shows a typical example of the identification of the incoming beam projectiles. Figure reprinted with permission from Adrich et al [27]. Copyright (2005) by the American Physical Society.

Standard image

The energy resolution typically achieved by this method ranges from 0.2 MeV close to the threshold to around 1 MeV around the giant-resonance region of about 12–15 MeV. The resolution is dominated by the determination of the excitation energy of the fragment, which is obtained by a measurement of the photon decays, i.e. the γ sum energy, and by the determination of the neutron kinetic energy. The instrumental response, however, might be rather complex due to deficiencies of the photon detection, or due to misidentification of the number of outgoing neutrons (or the fragment mass). This might lead to a distortion of the reconstructed spectrum which is usually accounted for by a complete simulation of the response when comparing data to theory. This procedure includes the simulation of the neutron decay of continuum states to (excited) fragment states with subsequent γ decay. Next-generation experiments aim at a calorimetric γ detection, a high-resolution tracking for heavy fragments, as well as high-efficient tracking of multi-neutron events in order to achieve a less complex energy response of the system.

4.2. Dipole strength distribution in neutron-rich oxygen isotopes and 26Ne

The first experiment measuring the dipole strength of unstable nuclei including the GDR region has been performed at GSI for neutron-rich oxygen isotopes 18−22O [26] and for the 1n channel for 23O [124]. From the measurement of the differential cross section dσ/dE* the photo-neutron cross section σ(γ,xn) has been deduced for 18,20,22O [26]. The deduced cross section for the stable isotope 18O is in good agreement with previous measurements using real photons. Figure 22 shows the results for the neutron-rich isotopes.

Figure 22.

Figure 22. Left panels: photo-neutron cross sections σ(γ,xn) for the unstable isotopes 20O (upper panel) and 22O (lower panel) as extracted from the measured electromagnetic excitation cross section (symbols) [26]. For comparison, the photo-neutron cross section for the stable 16O [127] is displayed as the solid curve in the upper panel. Right panels: evolution with neutron excess N − Z of integrated (up to 10 MeV excitation energy) strength $S_{\exp }$ in units of the Thomas–Reiche–Kuhn (TRK) sum rule STRK (upper panel) and of the cluster-sum-rule Sclus (lower panel) for oxygen isotopes. For the stable isotope 18O, experimental values from real-photon absorption experiments are given by the circle [125] and open square [126]. The values extracted from the electromagnetic-excitation experiment [26] are given by the filled squares. The data are compared with shell-model calculations of Sagawa and Suzuki [128] (solid curve in the lower left frame and stars in the right two frames) and the quasi-particle RPA calculation of Colò and Bortignon [129] (filled circles). Figure reprinted from Aumann [130]. With kind permission from Springer Science and Business Media.

Standard image

For comparison, the photo-absorption cross section for 16O measured with real photons [127] is overlaid on the spectrum for 20O in the upper left frame. In contrast to 16O, where the dipole strength is concentrated in the giant-resonance region around 24 MeV, the strength is much more fragmented for the neutron-rich isotopes. The low-lying dipole strength extends down to the neutron separation threshold around 8 MeV. The redistribution of the strength is reproduced by the shell-model calculation of Sagawa and Suzuki [128], as shown by the solid curve for 22O in the lower left frame. The experimental finding is also reproduced in relativistic RPA calculations by Vretenar et al [24]. In particular, the relativistic RPA not only reproduces the strong fragmentation, but also exhibits a low-lying peak in the dipole response of 22O at around 9 MeV, very close to the experimental finding, see figure 22. Vretenar et al [24] discuss the collectivity of the low-lying strength and conclude that the dipole transitions in the neutron-rich oxygen isotopes mainly result from single-particle transitions. Although the transition strength of the low-lying peak shows a characteristic shape for a neutron-skin oscillation as anticipated for the pygmy resonance. From the same theoretical calculations, the authors also concluded that the low-lying strength in heavier neutron-rich nuclei like Ni or Sn isotopes develops a more collective character. We will come back to the experimental results for the heavier nuclei later in this section.

The right panels of figure 22 show the integrated strength up to 10 MeV as a function of neutron excess in units of the energy-weighted sum rule (upper frame) and the energy-weighted cluster sum rule (lower frame) assuming a 16O core. About 5–10% of the TRK sum rule exhaustion is observed below 10 MeV corresponding to a large fraction of the cluster sum rule of 40–80% with a decreasing trend for the heavier isotopes. The magnitude and trend are well reproduced by the shell model [128] (stars) as well as the quasi-particle RPA calculation by Colò and Bortignon [129] (filled circles). It is interesting to note that if the strength is integrated up to 15 MeV, still in a region where no strength is found in 16O, the experimentally observed strength exceeds the cluster sum rule implying that also core excitations play a role in that energy region for the neutron-rich isotopes [26].

The experiments just discussed have measured the strength only above the neutron separation threshold. In an experiment at MSU, Tryggestad et al [131, 132] have measured the strength also below threshold for 18O and 20O by detecting the γ decays after Coulomb excitation at 100 MeV per nucleon. The measured γ spectrum is shown in the left part of figure 23, showing significant E1 strength in the region between 5 and 8 MeV, not present in the stable isotopes. The right part of the figure shows the extracted B(E1) and B(E2) strengths. The integrated value for the dipole strength below threshold is, however, much smaller than the one in the low-energy part above the threshold as discussed above.

Figure 23.

Figure 23. Left panel: Doppler-corrected γ energy spectrum measured after electromagnetic excitation of 100 MeV per nucleon 20O projectiles on a lead target [131, 132]. Right panels: dipole and quadrupole strength distributions below the neutron separation threshold as extracted from the γ spectrum (left frame). Figures reprinted with permission from Tryggestad et al [131]. Copyright (2002) by the American Physical Society.

Standard image

The low-energy dipole response has been studied in RIKEN for the neutron-rich isotope 26Ne by Gibelin et al [133] using Coulomb excitation at 58 MeV per nucleon on a Pb target. One neutron and γ-rays from the decay of the excited nucleus were measured in coincidence for invariant-mass analysis. The excitation-energy spectrum for the 1n channel is shown in the upper part of figure 24.

Figure 24.

Figure 24. Excitation-energy spectrum for the 1n-decay channel of 26Ne excited on a Pb target (upper frame) and an Al target (lower frame) [133]. The shaded area indicates a Lorentzian fit to the data. The one- and two-neutron separation thresholds are indicated. Figure reprinted with permission from Gibelin et al [133]. Copyright (2008) by the American Physical Society.

Standard image

The lower frame shows the cross section obtained with the Al target representing mainly nuclear excitations. The spectra show a peak-like structure at an excitation energy of around 9 MeV, rather close to the low-lying dipole structure predicted by Cao and Ma [134] at 8.5 MeV. Also the extracted integrated E1 strength, which exhausts 4.9(1.6)% of the Thomas–Reiche–Kuhn (TRK) sum rule, is in agreement with theoretical predictions. The analysis of microscopical calculations suggests that the low-lying structure mainly stems from single-particle transitions of neutrons occupying the least bound orbits in 26Ne (see discussion in [133] and references therein), similar to the interpretation discussed for the oxygen isotopes above. In the same experiment, the decay to the final fragment states has been analyzed as well. An interesting and surprising finding is that the excited 26Ne decays exclusively to excited states of 25Ne rather than to the ground state as predicted by the theoretical studies mentioned above.

4.3. Giant and PDRs in medium-heavy neutron-rich Ni and Sn isotopes

The low-energy dipole strength in light neutron-rich nuclei, as discussed above, has been interpreted theoretically as mainly arising from single-particle transitions rather than showing a collective nature as suggested by a macroscopic picture of the pygmy resonance of a collective vibration of excess neutrons versus a core nucleus. The onset of collectivity in the low-energy part of the dipole excitation spectrum for heavier nuclei has been discussed by Vretenar et al [24] in a relativistic RPA framework. They have shown for neutron-rich Ni and Sn isotopes that the low-energy peaks in the spectrum have a very different and more complicated character compared to the light nuclei involving many neutron p–h configurations. Vretenar et al concluded that the low-energy response in these nuclei indeed show the characteristics of a collective neutron vibration against the inert core composed of an equal number of neutrons and protons [24].

The first experimental evidence for a resonance-like structure in the low-energy dipole response was found for the neutron-rich isotopes 130,132Sn at GSI [27]. The experiment used secondary beams produced by fission of a primary uranium beam. The secondary beams impinged on a Pb target with energies around 500 MeV per nucleon. The differential cross section dσ/dE* for one-to-three neutron decay after electromagnetic excitation has been extracted and is shown for 130,132Sn in the left frames of the upper part of figure 25.

Figure 25.

Figure 25. Upper panel: cross section for electromagnetic excitation of Sn isotopes on Pb at around 500 MeV per nucleon (left frames) and corresponding photo-neutron cross section (right frames) [27]. The solid curve is a fit to the data containing a Lorentzian plus a Gaussian for the GDR (dash-dotted) and pygmy (dotted) contributions, respectively. Lower panel: pygmy dipole strength distribution dB(E1)/dE obtained for unstable Sn and Sb isotopes [30] with odd neutron number (upper row) and with even neutron number (bottom row). The solid lines show the results for 130,132Sn from the RQRPA calculation with a particular choice of the DD–ME interaction. Figures reprinted with permission from Adrich et al [27] and Klimkiewicz et al [30]. Copyright (2005) and (2007), respectively, by the American Physical Society.

Standard image

The right frames show the corresponding photo-neutron cross section, which can be directly compared with the photo-absorption cross section measured for stable nuclei. Evidently, the Lorentzian shape of the cross section for the GDR cannot account for the additional strength observed at low energy. The observed two-hump structure in the cross section is fitted by a Lorentzian for the GDR (dash-dotted curve) and a Gaussian accounting for the low-energy bump (dotted curve) which gives in total a good description of the measured spectrum (solid curve). The fit takes into account the experimental response including the distortions of the energy response due to gamma detection, misidentification of the number of neutrons or fragment mass, as well as all resolutions. The resulting parameters for the GDR for position width and sum rule exhaustion are in agreement with the expectation of the systematics of giant resonances collected for stable nuclei. Thus, although within the still large error bars, no systematic change of the frequency of the GDR mode in neutron-rich nuclei was observed. For the low-energy peak located at around 10 MeV, which exhausts around 5% of the energy-weighted sum rule, a lower limit on the width of 3.4 (2.5) MeV has been extracted for 130Sn (132Sn). The extracted strength agrees well with several theoretical predictions. It should be noted, however, that the predictions depend strongly on the effective interactions used, in particular on the symmetry energy as will be discussed below. The position of the pygmy mode is predicted in relativistic RPA calculations around 8–9 MeV, for some Skyrme forces around 10 MeV, i.e. closer to the experimental finding. For a comparison and discussion of the different theoretical results and models, see the review article of Paar et al [25].

The same collaboration has published data on low-lying strength in a couple of nuclei around 130,132Sn [30] including nuclei with odd neutron number. The results after subtracting the GDR contribution and after conversion into dipole strength B(E1) is displayed in the lower part of figure 25. In all nuclei, strength of a similar magnitude is observed located between 6 and 12 MeV. It is interesting to note that the odd nuclei have significant lower neutron separation thresholds compared to even ones extending the sensitivity of the experiment to lower excitation energies (the experiment relied on the measurement in coincidence with at least one neutron). A clear odd–even staggering in the centroid is visible. The interpretation of this effect is not clear. It could be related to an odd–even staggering of the position of the pygmy strength, or it could be explained by missing strength for the even isotopes due to the higher threshold. It is evidently mandatory, that experiments with an extended detection scheme including the measurement of γ decays below the threshold have to be carried out in order to understand this behavior. The integrated strength in excess to the GDR tail for neutron-rich tin isotopes is significantly larger than the one found in stable nuclei in the same mass range. It has to be noted, however, that the strength extracted from (γ,γ') measurements at stable nuclei provides often a lower limit only since the measurement is sensitive only to the direct γ decay back to the ground state, which might be a small fraction depending on structure and level density. For the measurements with neutron-rich tin isotopes discussed above, on the other hand, additional non-observed strength below the threshold cannot be excluded. New more complete experiments are needed for both stable and unstable nuclei for clarification and for a proper comparison of the measurements for stable and unstable nuclei. For a more complete discussion including a comparison to the experimental findings with stable nuclei, see the recent review by Savran et al [22].

Another method based on heavy-ion-induced electromagnetic excitation has been employed by the RISING collaboration. In an experiment performed at GSI, the γ decay of 68Ni projectiles at 600 MeV per nucleon has been measured after excitation on gold target nuclei [135]. As can be seen in the right part of figure 26, a peak-like structure can be observed in the Doppler corrected γ spectrum above a background stemming mainly from statistical decay of the GDR excited in the target nuclei or in the projectile.

Figure 26.

Figure 26. Left: cross section for electromagnetic excitation and photon decay for 68Ni at 600 MeV per nucleon on an Au target [135]. The Doppler corrected γ spectrum has been measured with BaF2 detectors at GSI at the FRS and is compared to statistical-model calculations for the decay after target excitation (dotted line) and projectile excitation (dashed line). The inset shows a comparison of a GEANT simulation for a γ-transition at 11 MeV. The right panel shows the contributions of the GDR and the PDR to the background.subtracted γ spectrum. Figures reprinted with permission from Wieland et al [135]. Copyright (2009) by the American Physical Society.

Standard image

The excess cross section is peaked at around 11 MeV, i.e. above the neutron separation threshold at around 7.8 MeV. The apparent width of the peak mainly reflects the experimental resolution. The inset shows a GEANT simulation assuming a γ transition at 11 MeV with a very narrow width. The position of the peak is close to the predicted value by the relativistic RPA calculations of Vretenar et al [24] of 9 MeV, and the relativistic quasiparticle RPA (RQRPA) of Cao and Ma [134]. The result of an extended RQRPA calculation including coupling to more complicated states (two-phonon relativistic quasiparticle time blocking approximation) of Litvinova et al [136] shows a centroid of the pygmy strength at slightly higher energy peaking at 10.3 MeV rather close to the experimental finding.

The integrated cross section of the peak cannot be directly translated into a B(E1) value for the excitation since the transition lies above the neutron threshold resulting in a rather small probability for a direct photon decay. Wieland et al [135] estimated this branching ratio assuming a level density to be of the order of 4%, which yields a dipole strength for the 11 MeV peak corresponding to about 5% of the energy-weighted sum rule. The level density used was significantly smaller than the systematic one for stable nuclei. If using the latter would result in a smaller decay branch and thus in a larger B(E1) value corresponding to 9% of the sum rule. A new experiment measuring the neutron decay after electromagnetic excitation of 68Ni has been carried out at the R3B-LAND setup at GSI [137]. Preliminary results shown on conferences show that low-lying dipole strength has been observed [137], which is located between 10 and 11 MeV and exhausting 4.1(1.9)% of the energy-weighted sum rule, which is in agreement with the RISING experiment. The final analysis will also allow for extracting the γ decay branching ratio for the pygmy resonance.

4.4. The dipole response of nuclei and the neutron skin

Since the low-lying dipole strength is related to an excitation of the less-well bound excess neutrons, it is clear that the low-lying dipole response should be related to the development of a neutron skin in medium–heavy neutron-rich nuclei. A quantitative analysis of this was first presented by Piekarewicz [28] who proposed that the systematics of the pygmy dipole mode may be used to constrain the density dependence of the symmetry energy. It has been pointed out by many authors, see for instance [138, 139], that the neutron-skin thickness calculated in mean-field models depends strongly on the density dependence of the symmetry energy (close to saturation density) incorporated into the effective interactions used. A similar dependence is found for the low-lying dipole strength. Klimkiewicz et al [30] used a relativistic interaction with a systematically changed symmetry-energy parameter a4 and found an almost linear correlation with the low-lying dipole strength and the neutron-skin thickness, see figure 27.

Figure 27.

Figure 27. Left panel: summed B(E1) strength exhausted by the low-energy peak relative to the strength exhausted by the GDR (upper frames) and neutron-skin thickness (lower frames) calculated in relativistic RPA as a function of the symmetry-energy parameter a4 used in the calculation for 130Sn (left frames) and 132Sn (right frames). The dashed lines in the upper frames indicate the experimental value for the strength of the 10 MeV peak resulting in an average symmetry parameter a4 = 32.0(1.8) MeV; this result corresponds to neutron-skin thicknesses as indicated in the lower frames. Right panel: parameter L describing the density dependence of the symmetry energy as deduced by various approaches (see [31]). The values deduced from the PDR by Carbone et al [31] are based on RPA calculations with a variety of different Skyrme forces. Figures reprinted with permission from Klimkiewicz et al [30] and Carbone et al [31]. Copyright (2007) and (2010), respectively, by the American Physical Society.

Standard image

In this procedure, all other parameters of the interaction have been refitted to properties of stable nuclei in order to ensure that always a well-calibrated force is used. From the comparison with the data on 130,132Sn discussed above, they deduced a value a4 = 32.0(1.8) MeV for the symmetry energy (upper panels of figure 27) and a corresponding neutron-skin thickness for 132Sn of 0.24(4) fm (see the lower panels of figure 27).

A similar analysis based on RPA calculations with different Skyrme effective forces has been performed by Carbone et al [31]. Comparing the theoretical calculations with the experimental results on 132Sn and 68Ni discussed above, they deduced a value for the slope parameter L, which is related to the density dependence of the symmetry energy at saturation density. The results are compared in figure 27 to the results obtained by other methods. The results are in good agreement with the values deduced using other approaches, as can be seen in figure 27. It should be noted, however, that the procedure is not model independent. It should also be noted that the experimental value for 68Ni assumed a γ decay branch to extract the B(E1) strength, see the discussion above. Another difficulty in the method is defining the PDR strength in theory and experiment which have to be compared. Depending on the theoretical calculation, the strength observed in addition to the GDR tail does not peak at the same excitation energy as in the calculation. It has been pointed out by Reinhard and Nazarewicz [29] that a better correlated and more robust observable would be the polarizability of the nucleus. Since the polarizability is the inversely energy-weighted integral of the E1 strength, the need of defining an energy window for the low-energy dipole strength is not necessary, while the value of the polarizability is still very sensitive to a redistribution of the dipole strength. Based on such calculations, Tamii et al [32] deduced a value for the neutron-skin thickness of 208Pb of 0.156(25) fm from a precise measurement of the polarizability of 208Pb deduced from the measured Coulomb excitation cross section in (p,p) scattering close to zero degree [32]. A value of 0.168(22) fm has been deduced by Piekarewicz et al [141] from the same experimental measurement.

Although the relative effect of the density dependence of the symmetry energy, or correspondingly the thickness of the neutron skin, is larger for the strength of the low-energy part of dipole response as compared to the polarizability, a less strong model dependence can be a big advantage. An advantage is here to study exotic nuclei as neutron-rich as possible to maximize the effect. The influence of different symmetry-energy parameters in the mean-field model on the calculated dipole response can clearly be seen in figure 28, where the dipole response calculated in relativistic RPA is shown for 68Ni [141].

Figure 28.

Figure 28. Dipole response of 68Ni calculated in a relativistic RPA approach using effective forces with different slope parameters for the symmetry energy [141] resulting in different values for the predicted neutron-skin thicknesses (corresponding values for 208Pb are indicated). The left panel shows the energy-weighted strength and its cumulative value in the inset, whereas the right panel shows the inverse energy-weighted strength. Figure reprinted with permission from Piekarewicz [140]. Copyright (2011) by the American Physical Society.

Standard image

The left panel of figure 28 shows the energy-weighted response for 68Ni calculated with effective forces using different values for the symmetry energy resulting in different calculated neutron-skin thicknesses as well as in a different shape of the response [141]. The corresponding value of the neutron-skin thickness calculated for 208Pb is indicated in the figure. The larger the neutron-skin thickness, the more the strength is shifted from high excitation energies to lower energy. The inset shows the running sum of the energy-weighted strength, i.e. the exhaustion of the energy-weighted dipole sum rule as a function of excitation energy. It is clearly seen that while the shape of the spectrum changes significantly, the curves converge if integrated above the GDR region, i.e. each calculation respects the sum rule. The change in the shape of the response is more drastically seen in the right frame, where the inversely weighted strength is plotted, i.e. the polarizability as a function of excitation energy. The running sum keeps diverging until a flat region is reached if integrated above the GDR region. Most noticeably, the polarizability calculated with the different effective forces differs by about 20% for a range of parameters shown. This effect is much bigger than in 208Pb, thus demonstrating the large sensitivity one might reach with radioactive beams. Even a measurement with moderate accuracy could yield a rather precise value for the neutron-skin thickness or the slope parameter of the symmetry energy.

5. Conclusion and outlook

The characteristic changes of the properties of nuclei when going from stable to neutron–proton asymmetric nuclei, such as the asymmetric Fermi energies for protons and neutrons and the development of neutron skins and haloes, manifest themselves in the electric dipole response of these nuclei. From an experimental point of view, such effects have been studied so far mainly for neutron-rich nuclei. An extreme case is the halo nuclei at the neutron-drip line for which a huge dipole-transition strength is found just above the neutron separation threshold. The decoupling of the loosely bound neutrons exhibiting an extended density distribution from the well-bound core nucleons with a normal density distribution is reflected in the decoupling of the dipole response of the halo, which is observed as low-energy threshold strength, and the response of the core, which we believe corresponds to the usual GDR. It has to be said, however, that the experimental determination of the full strength function including the GDR has not been possible so far for a halo nucleus. The characteristic features of the soft E1 excitation related to the halo are meanwhile understood quantitatively for the one-neutron halo nuclei and used as a tool to study the single-particle structure of the halo wave function, such as its spatial extension, its single-particle configurations, and the associated spectroscopic factors. For two- or more-neutron haloes, the dipole response contains additional information on the spatial correlations among the halo neutrons, which in turn can be investigated by Coulomb-breakup experiments. The method of Coulomb breakup is thus an ideal tool to study heavier drip-line nuclei in the future, as they will become accessible at the new fragmentation-beam facilities such as RIBF at RIKEN, FRIB at MSU and FAIR at Darmstadt.

Measurements of the electric dipole response of neutron-rich non-halo nuclei, and in particular of heavier nuclei, are still rather scarce. For the oxygen isotopic chain, a strong fragmentation of the dipole strength has been found for the neutron-rich isotopes, but a clear separation of giant-dipole excitations and low-lying excitations of excess neutrons is not evident. For the heavier neutron-rich nuclei around 132Sn and 68Ni, evidence for a low-lying resonance has been observed, which is understood by microscopical models as a vibration of excess neutrons versus a core. A systematic investigation is still lacking, and even a comparison with the results from stable nuclei is hampered by different experimental sensitivities. Next-generation experiments are already on the way both for stable as well as for radioactive beams at the different facilities. In particular, the newly developed instrumentation such as SAMURAI at RIKEN and R3B at GSI and later FAIR will provide the basis for improved and more complete measurements. The new facilities will not only allow to study nuclei with larger neutron excess, but also in a wider mass range including neutron-rich lead isotopes which could be studied at R3B at FAIR. In addition, other experimental techniques have now become available that will allow investigation of the dipole response of exotic nuclei using different probes, such as the α scattering to study the isoscalar dipole response. Finally, with a measurement of the full strength function below and above the threshold including the GDR it will be possible to extract the polarizability of neutron-rich nuclei yielding important information on the neutron-skin thickness and the isospin-asymmetric part of the equation of state of nuclear matter.

In this article, we have concentrated on experimental results and developments. Progress in understanding the dipole response of neutron-rich nuclei, however, has been made in conjunction with theoretical efforts in this field over the last two decades. The fact that both theory and experiment are intensively developing in this field of research makes us confident that progress on answering the many still open questions related to the dipole response of exotic nuclei will be made in the near future.

Acknowledgments

We would like to thank all colleagues who contributed to the experiments discussed here. In particular, we thank our colleagues who have allowed us to use the figures from experiments we were not involved in. TA is much indebted to the members of the LAND-, GSI-Halo-, and R3B collaborations and especially to his long-term GSI colleagues K Boretzky, L Chulkov, H Emling, D Rossi and H Simon. TN would like to thank the collaborators in the RIPS collaboration for the breakup experiments of halo nuclei, and the ZDS- and SAMURAI collaborations at RIBF at RIKEN, and in particular, Y Kondo, H Satou, T Kobayashi, T Motobayashi and K Yoneda. Both authors would like to thank the late P G Hansen for giving us invaluable inspiration and insights about physics of exotic nuclei. We gratefully acknowledge many valuable discussions with B Jonson, M Ishihara and I Tanihata. TA acknowledges support by the Helmholtz International Center for FAIR, GSI and the BMBF project 06DA7047I. TN has been supported in part by a Grant-in-Aid for Scientific Research (B) (no. 22340053), and that on Innovative Areas (no. 24115005) from the Ministry of Education, Culture, Sports, Science and Technology (MEXT, Japan).

Footnotes

  • The term 'Coulomb breakup' is equivalent to Coulomb dissociation. The term EMD implies a possibility of including magnetic excitation, although otherwise it is equivalent to Coulomb breakup.

  • Here the 6° for the Pb target and the 12° for the C target correspond to the whole acceptance. This difference corresponds to that of the c.m. angles which is about double for the C target for the same laboratory angle.

Please wait… references are loading.
10.1088/0031-8949/2013/T152/014012